Optoelectronic Organic-Inorganic Semiconductor Heterojunctions 9781000325751, 100032575X

Optoelectronic Organic-Inorganic Semiconductor Heterojunctions summarizes advances in the development of organic-inorgan

291 78 11MB

English Pages 364 [382] Year 2021

Report DMCA / Copyright

DOWNLOAD PDF FILE

Table of contents :
Cover
Half Title
Title Page
Copyright Page
Table of Contents
Preface
Editor Biography
List of Contibutors
Chapter 1: Introduction to Organic–Inorganic Heterojunction
References
Chapter 2: Energy-Level Alignment at Organic–Inorganic Heterojunctions
2.1 Introduction
2.2 Interface Formation between Organic and Inorganic Semiconductors: 6P on ZnO
2.3 Work Function Tuning of ZnO via Dipole Bearing Self-assembled Monolayers
2.4 Work Function Tuning with Electron Donor and Acceptor Molecules
2.5 Fingerprint of Ground-State Charge Transfer in the Optical Spectra of ZnO-Acceptor Interfaces
2.6 Organic–Inorganic Semiconductor pn-Junction
2.7 Energy-Level Tuned Organic–Inorganic Heterojunctions for Light-Emitting Applications
Acknowledgements
References
Chapter 3: Molecular Layer Deposition of Organic–Inorganic Hybrid Materials
3.1 Introduction
3.2 The Basics of MLD
3.2.1 Surface Chemistry
3.2.2 Growth Characteristics
3.3 MLD Processes for Organic–Inorganic Hybrid Metalcones
3.3.1 Alucones
3.3.1.1 Homobifunctional Organic Precursors
3.3.1.2 Heterobifunctional Organic Precursors
3.3.2 Titanicones
3.3.3 Zincones
3.3.4 Other Metalcones
3.4 Other Hybrid Materials
3.4.1 Luminescent Hybrid Materials
3.4.2 Metal-Organic Frameworks (MOFs)
3.4.3 Energy-Storage Materials
3.4.4 Organic Magnets
3.4.5 Complex MLD Processes
3.4.6 Organic–Inorganic Hybrid Nanolaminates by MLD and ALD
3.5 Conclusions
Acknowledgements
References
Chapter 4: Scanning Tunneling Microscope and Spectroscope on Organic–Inorganic Material Heterojunction
4.1 Introduction
4.2 Band Mapping; across a PN-Junction in a Nanorod
4.2.1 Nanorods and Junctions Characterization: Tunneling Current and Density of States
4.2.2 Parallel PN-Junctions across Nanowires via One-Step Ex Situ Doping
4.3 Interfacial Band Mapping across Vertically Phased Separated Polymer/Fullerene Hybrid Solar Cells
4.4 Organic–Inorganic Hybrid Heterojunction
4.4.1 Photocarrier Generations and Band Alignments at Perovskite/PbI 2 Heterointerfaces
4.4.2 Photocarrier Generations of Perovskites during Illumination
4.4.3 Band Alignments of Perovskites during Illumination
4.4.4 PbI 2 Layer Thickness Dependence of ΔED
4.5 Outlook and Upcoming Challenges
References
Chapter 5: Organic–Inorganic Semiconducting Nanomaterial Heterojunctions
5.1 Overview
5.2 Heterojunction of Cd-based Inorganic Semiconductor
5.3 Heterojunction Nanodots of Zn-based Inorganic Semiconductors
5.4 Heterojunction of Ti-based Inorganic Semiconductors
5.5 Heterojunction of Si-based Inorganic Semiconductors
5.6 Heterojunction of Perovskite-based Inorganic Semiconductors
5.7 Heterojunction of Ag-based Inorganic Semiconductors
5.8 Heterojunction of Bi-based Inorganic Semiconductors
5.9 Heterojunction of Pb-based inorganic semiconductors
5.10 Heterojunction of Other Metal-based Inorganic Semiconductors
5.11 Conclusions
Abbreviations and Acronyms
References
Chapter 6: Organic–Inorganic Heterojunction Nanowires
6.1 Introduction: Background and Driving Forces
6.2 The Synthetic Methods of Organic–Inorganic Heterojunction Nanowires
6.2.1 Solution Phase Method
6.2.2 Template Method Combined with Electrochemical Polymerization
6.2.2.1 Template Method Combined with Pressure Injection
6.2.3 Vapor–Liquid–Solid Method
6.3 The Applications of Organic–Inorganic Heterojunction Nanowires
6.3.1 Field Emission
6.3.2 Diode Rectification
6.3.3 Solar Cells
6.3.4 Photoelectric Detection
6.3.5 Logic Gates
6.4 Summary and Perspective
References
Chapter 7: Electroluminescence of Organic Molecular Junction in Scanning Tunneling Microscope
7.1 Introduction: Molecular Junctions and Devices
7.2 Transport Mechanism in Molecular Junctions
7.2.1 Coherent Transport
7.2.2 Incoherent Transport
7.3 Optical Properties of Molecular Junctions
7.4 Special Phenomena: Hot Luminescence and Upconversion
7.4.1 Hot Luminescence
7.4.2 Upconversion Electroluminescence
7.5 Summary and Outlook
References
Chapter 8: Recent Research Progress on Organic–Inorganic Hybrid Solar Cells
8.1 Introduction
8.2 ZnO organic Hybrid Solar Cells
8.2.1 ZnO-NP Organic Hybrid Solar Cells
8.2.2 Modified ZnO Organic Hybrid Solar Cells
8.3 TiO 2 organic Hybrid Solar Cells
8.3.1 TiO 2 -NP Organic Hybrid Solar Cells
8.3.2 Modified TiO 2 Organic Hybrid Solar Cells
8.4 ZnO/TiO 2 Organic Hybrid Solar Cells
8.5 New Type Organic–Inorganic Solar Cells based on All Chl Derivative
References
Chapter 9: Nanogenerators Based on Organic–Inorganic Heterojunction Materials
9.1 Introduction
9.2 Fundamentals of Nanogenerator
9.3 Piezoelectric Nanogenerators Based on Organic–Inorganic Hybrid Nanomaterial
9.3.1 Basic Concept of PENGs and Its Operating Principle
9.3.2 Material Design Criteria and Techniques for Performance Enhancement
9.3.3 InN Nanowire-Based High-Performance PENGs
9.3.4 1D/2D ZnO Nanostructure-Based PENGs
9.4 Triboelectric Nanogenerators Based on Organic–Inorganic Hybrid Nanomaterial
9.4.1 Basic Concept of TENGs and Its Operating Principle
9.4.2 Material Design Criteria and Techniques for Performance Enhancement
9.4.3 High-Performance TENGs
9.5 Hybrid nanogenerators Based on Organic–Inorganic Hybrid Nanomaterial
9.5.1 Basic Concept of HNGs and Its Operating Principle
9.5.2 Various Approaches Taken to Design High-Performance HNGs
9.5.2.1 Cascade-Type Hybrid Nanogenerator
9.5.2.2 Organic–Inorganic Hybrid NG
9.6 Conclusion
Acknowledgments
References
Chapter 10: Organic–Inorganic Semiconductor Heterojunctions for Hybrid Light-Emitting Diodes
10.1 Introduction
10.2 Basic Introduction to White LEDs
10.2.1 III-Nitride Semiconductors and Inorganic LEDs
10.2.2 Colorimetry, Radiometry, Photometry, and Efficacy
10.2.3 White Light Generation
10.2.4 Use of Phosphors in White LEDs
10.2.5 White Organic LEDs
10.3 Chemistry
10.4 Light-Emitting Polymers
10.4.1 Introduction
10.4.2 Polymers in Hybrid White LEDs
10.5 Luminescent Small Molecules
10.5.1 Introduction
10.5.2 Manipulation of the Chemical Structure and Effect on Optical Properties
10.5.2.1 Introduction to BODIPY
10.5.2.2 Toward White Light: Yellow Emission from Oligofluorene-BODIPY Oligomers
10.5.2.3 Toward White Light: Blue Light Absorption for White LEDs
10.5.2.4 Toward White Light: Nanorod Encapsulation
10.5.3 Toward White Light: Deposition and Encapsulation
10.5.4 White Light Device Efficiency and Efficacy
10.5.5 White Light Degradation and Lifetime
10.5.6 Next-Generation White-Emitting LEDs with Improved Efficacy
10.5.7 Metal–Organic Frameworks
10.6 Summary
Acknowledgments
References
Chapter 11: Organic–Inorganic Semiconductor Heterojunctions for Resistive Switching Memories
11.1 Introduction
11.2 Organic–Inorganic Semiconductor Heterojunctions for WORM Memory Devices
11.3 Organic–Inorganic Semiconductor Heterojunctions for Unipolar Memory Devices
11.4 Organic–Inorganic Semiconductor Heterojunctions for Bipolar Memory Devices
11.5 Challenges and Prospects
Acknowledgments
References
Chapter 12: Optoelectronic Sensors for Health Monitoring
12.1 Introduction
12.2 Sensing Mechanisms and Materials
12.2.1 Active Sensing Components
12.2.2 Supporting Substrates and Fabrication Methods
12.2.2.1 Photolithography
12.2.2.2 Functional Printing of Electronic Sensing Devices
12.2.2.3 Fabrication of Optical Sensor Arrays
12.3 Clinical Applications of Optoelectronic Sensors
12.3.1 Physical Index Monitoring
12.3.2 Human Movement Monitoring
12.3.3 Chemical Index Monitoring
12.3.3.1 Glucose
12.3.3.2 Pathogens
12.3.3.3 Cancer and Other Disease Biomarkers
12.4 Conclusions
References
Chapter 13: Organic–Inorganic Semiconductor Heterojunction Photocatalysts
13.1 Introduction
13.2 Photocatalysts
13.2.1 Inorganic Photocatalyst
13.2.1.1 Metal Oxide
13.2.1.2 Sulfide
13.2.1.3 Solid Solution
13.2.1.4 Perovskite
13.2.2 Organic Photocatalyst
13.2.2.1 C 3 N 4 -Based Materials
13.2.2.2 MOF
13.2.2.3 PDI
13.2.3 Organic–Inorganic Heterojunction Photocatalyst
13.2.3.1 C 3 N 4 -Inorganic Heterojunction
13.2.3.2 MOFs-Inorganic Heterojunction
13.2.3.3 PDI-Inorganic Heterojunction
13.3 Mechanism of the Photocatalysis for Heterojunctions
13.3.1 p-n Heterojunction
13.3.2 Type I and Type II Heterojunctions
13.3.3 Z-Scheme Heterojunction
13.4 Conclusion and Outlook
References
Index
A
B
C
D
E
F
G
H
I
J
K
L
M
N
O
P
Q
R
S
T
U
V
W
X
Y
Z
Recommend Papers

Optoelectronic Organic-Inorganic Semiconductor Heterojunctions
 9781000325751, 100032575X

  • 0 0 0
  • Like this paper and download? You can publish your own PDF file online for free in a few minutes! Sign Up
File loading please wait...
Citation preview

Optoelectronic Organic–Inorganic Semiconductor Heterojunctions

Optoelectronic Organic–Inorganic Semiconductor Heterojunctions

Ye Zhou

First edition published 2021 by CRC Press 6000 Broken Sound Parkway NW, Suite 300, Boca Raton, FL 33487-2742 and by CRC Press 2 Park Square, Milton Park, Abingdon, Oxon, OX14 4RN © 2021 Taylor & Francis Group, LLC CRC Press is an imprint of Taylor & Francis Group, LLC The right of Ye Zhou to be identified as the author of the editorial material, and of the authors for their individual chapters, has been asserted in accordance with sections 77 and 78 of the Copyright, Designs and Patents Act 1988. Reasonable efforts have been made to publish reliable data and information, but the author and publisher cannot assume responsibility for the validity of all materials or the consequences of their use. The authors and publishers have attempted to trace the copyright holders of all material reproduced in this publication and apologize to copyright holders if permission to publish in this form has not been obtained. If any copyright material has not been acknowledged, please write and let us know so we may rectify in any future reprint. Except as permitted under U.S. Copyright Law, no part of this book may be reprinted, reproduced, transmitted, or utilized in any form by any electronic, mechanical, or other means, now known or hereafter invented, including photocopying, microfilming, and recording, or in any information storage or retrieval system, without written permission from the publishers. For permission to photocopy or use material electronically from this work, access www.copyright.com or contact the Copyright Clearance Center, Inc. (CCC), 222 Rosewood Drive, Danvers, MA 01923, 978750-8400. For works that are not available on CCC, please contact [email protected] Trademark notice: Product or corporate names may be trademarks or registered trademarks and are used only for identification and explanation without intent to infringe. Library of Congress Cataloging-in-Publication Data Names: Zhou, Ye (Semiconductor engineer), editor. Title: Optoelectronic organic-inorganic semiconductor heterojunctions / [edited] by Ye Zhou. Description: First edition. | Boca Raton, FL : CRC Press/Taylor & Francis Group, LLC, 2021. | Includes bibliographical references and index. | Summary: “The book summarizes advances in development of organic-inorganic semiconductor heterojunctions, challenges and possible solutions for material/device design, and prospects for commercial applications. It introduces the concept and basic mechanism of semiconductor heterojunctions. It describes a series of organic-inorganic semiconductor heterojunctions with desirable electrical and optical properties for optoelectronic devices. Typical devices such as solar cells, photo-detectors and optoelectronic memories are discussed. Materials, device challenges, and strategies are discussed to promote the commercial translation of semiconductor heterojunction based optoelectronic devices” -- Provided by publisher. Identifiers: LCCN 2020043631 (print) | LCCN 2020043632 (ebook) | ISBN 9780367342128 (hardback) | ISBN 9780367348175 (ebook) Subjects: LCSH: Heterojunctions. | Optoelectronic devices. | Organic semiconductors. Classification: LCC TK7874.53 .O68 2021 (print) | LCC TK7874.53 (ebook) | DDC 621.3815/2--dc23 LC record available at https://lccn.loc.gov/2020043631 LC ebook record available at https://lccn.loc.gov/2020043632 ISBN: 978-0-367-34212-8 (hbk) ISBN: 978-0-367-34817-5 (ebk) Typeset in Times by SPi Global, India

Contents Preface����������������������������������������������������������������������������������������������������������������������vii Editor Biography�������������������������������������������������������������������������������������������������������ix Contibutors����������������������������������������������������������������������������������������������������������������xi Chapter 1

Introduction to Organic–Inorganic Heterojunction............................. 1 Kui Zhou and Ye Zhou

Chapter 2

Energy-Level Alignment at Organic–Inorganic Heterojunctions........ 9 Sylke Blumstengel and Norbert Koch

Chapter 3

Molecular Layer Deposition of Organic–Inorganic Hybrid Materials................................................................................ 37 Xiangbo Meng

Chapter 4

Scanning Tunneling Microscope and Spectroscope on Organic–Inorganic Material Heterojunction................................. 71 Sadaf Bashir Khan and Shern Long Lee

Chapter 5

Organic-Inorganic Semiconducting Nanomaterial Heterojunctions................................................................................ 101 Jie Guan, Ziwei Wang, Yuan-Cheng Zhu, Wei-Wei Zhao, and Qin Xu

Chapter 6

Organic–Inorganic Heterojunction Nanowires................................ 127 Yuan Yao and Yanbing Guo

Chapter 7

Electroluminescence of Organic Molecular Junction in Scanning Tunneling Microscope................................................. 147 Xiaoguang Li

Chapter 8

Recent Research Progress on Organic–Inorganic Hybrid Solar Cells........................................................................... 165 Wenjie Zhao, Na Li, Xin Jin, Shengnan Duan, Baoning Wang, Aijun Li, and Xiao-Feng Wang

v

viContents

Chapter 9

Nanogenerators Based on Organic–Inorganic Heterojunction Materials................................................................. 197 Md Masud Rana, Asif Abdullah Khan, and Dayan Ban

Chapter 10 Organic–Inorganic Semiconductor Heterojunctions for Hybrid Light-Emitting Diodes................................................... 231 J. Bruckbauer and N. J. Findlay

Chapter 11 Organic–Inorganic Semiconductor Heterojunctions for Resistive Switching Memories................................................... 267 Shuang Gao and Run-Wei Li

Chapter 12 Optoelectronic Sensors for Health Monitoring................................ 287 Zheng Li

Chapter 13 Organic–Inorganic Semiconductor Heterojunction Photocatalysts......................................................... 315 Tao Lv, Zhengyuan Jin, Luhong Zhang, and Yu-Jia Zeng

Index..�������������������������������������������������������������������������������������������������������������������� 351

Preface A semiconductor heterojunction is the interface between two layers or regions of different semiconductor materials. When combining organic semiconductor with inorganic semiconductor, a hybrid heterojunction can be formed, either in planar structure or bulk structure. Organic–inorganic heterojunctions are promising candidates for functional electronic devices owing to the expected co-activation between organic and inorganic components. The hybrid system is beneficial from both the advantages of organic semiconductors and inorganic counterpart. Optoelectronic semiconductor organic–inorganic heterojunctions have attracted intense research interests in the past decade. If the two semiconductors are combined through a reasonable interface design, the strengths can be enhanced. It is expected to obtain better photoelectric performance with proper design of the heterojunction materials. These kinds of heterojunctions have great potential to be applied in transistors, light-emitting diodes, sensors, solar cells, and memories. Scientists have further explored the one-dimensional and zero-dimensional electronic behaviors in heterogeneous structures. It is expected that new phenomena will be discovered in the future, and more novel heterogeneous structural components will emerge. In this book, we have introduced the concept and basic mechanism of semiconductor heterojunctions. We investigate a series of organic–inorganic semiconductor heterojunctions with desirable electrical and optical properties for optoelectronic devices. The typical devices such as solar cells, photo-detectors, and optoelectronic memories are discussed. Materials and device challenges as well as possible strategies are also discussed to promote the commercial translation of these semiconductor ­heterojunctions-based optoelectronic devices. I would like to acknowledge all the fellow authors who have contributed in this book. I also want to express my gratitude to Gabrielle Vernachio, Allison Shatkin, Lara S. Loes, and Camilla Michael at CRC Press/Taylor & Francis and Karthik Thiruvengadam at SPi Global, for all the help during the book editorial process, and for the excellent experience of working with them. Specially, I want to thank all the readers for their interest in our book. Our aim is to give a comprehensive, critical, and up-todate book here. I hope that this book can be useful as a reference guide for researchers and students who work in the field of organic–­inorganic heterojunctions and related electronic devices. Ye Zhou

vii

Editor Biography Prof. Ye Zhou is an IAS Fellow and group leader in the Institute for Advanced Study, Shenzhen University. He received his B.S. (2008) in Electronic Science and Engineering from Nanjing University, M.S. (2009) in Electronic Engineering from Hong Kong University of Science and Technology, and Ph.D. (2013) in Physics and Materials Science from City University of Hong Kong. His research interests include flexible and printed electronics, nano-materials, nano-composite materials, and nano-scale devices for technological applications such as logic circuits, memories, photonics and sensors. He has edited 4 books, 3 USA patents, 11 China patents, and over 120 SCI papers in journals such as Science, Nature Electronics, Nature Communications, Chemical Reviews, Advanced Materials, etc. More than 40 of his works have been highlighted as cover pages or frontispieces. These published papers have been extensively accessed and cited by prestigious journals. He is the Associate Editor of STAM, Applied Nanoscience and IEEE Access, and sits at Editorial/ Community Board of Materials Horizons, Multifunctional Materials, Chemistry, PLOS ONE, and Chemistry Proceedings.

ix

Contibutors Dayan Ban Department of Electrical and Computer Engineering Waterloo Institute for Nanotechnology, University of Waterloo Waterloo, Ontario, Canada Sylke Blumstengel Institut für Physik, Institut für Chemie & IRIS Adlershof Humboldt-Universität zu Berlin Berlin, Germany J. Bruckbauer Department of Physics SUPA, University of Strathclyde Glasgow, Scotland, UK Shengnan Duan Key Laboratory of Physics and Technology for Advanced Batteries (Ministry of Education) College of Physics, Jilin University Changchun, P. R. China N. J. Findlay WestCHEM, School of Chemistry University of Glasgow Glasgow, Scotland, UK Shuang Gao CAS Key Laboratory of Magnetic Materials and Devices & Zhejiang Province Key Laboratory of Magnetic Materials and Application Technology Ningbo Institute of Materials Technology and Engineering, Chinese Academy of Sciences Ningbo, Zhejiang, China

Jie Guan School of Chemistry and Chemical Engineering Yangzhou University Yangzhou, Jiangsu, China Yanbing Guo Central China Normal University Wuhan, Hubei, China Zhengyuan Jin College of Physics and Optoelectronic Engineering Shenzhen University Shenzhen, China Xin Jin Key Laboratory of Physics and Technology for Advanced Batteries (Ministry of Education) College of Physics, Jilin University Changchun, P. R. China Sadaf Bashir Khan Institute for Advanced Study Shenzhen University Shenzhen, Guangdong, China Asif Abdullah Khan Department of Electrical and Computer Engineering Waterloo Institute for Nanotechnology, University of Waterloo Waterloo, Ontario, Canada Norbert Koch Institut für Physik & IRIS Adlershof Humboldt-Universität zu Berlin Berlin, Germany Helmholtz-Zentrum Berlin für Materialien und Energie Berlin, Germany

xi

xiiContibutors

Shern Long Lee Institute for Advanced Study Shenzhen University Shenzhen, Guangdong, China

Xiangbo Meng Department of Mechanical Engineering University of Arkansas Fayetteville, Arkansas, USA

Run-Wei Li CAS Key Laboratory of Magnetic Materials and Devices & Zhejiang Province Key Laboratory of Magnetic Materials and Application Technology Ningbo Institute of Materials Technology and Engineering, Chinese Academy of Sciences Ningbo, Zhejiang, China

Md Masud Rana Department of Electrical and Computer Engineering Waterloo Institute for Nanotechnology, University of Waterloo Waterloo, Ontario, Canada

Xiaoguang Li Institute for Advanced Study Shenzhen University Shenzhen, Guangdong, China Zheng Li Institute for Advanced Study Shenzhen University Shenzhen, Guangdong, China Na Li Key Laboratory of Physics and Technology for Advanced Batteries (Ministry of Education) College of Physics, Jilin University Changchun, P. R. China Aijun Li Key Laboratory of Physics and Technology for Advanced Batteries (Ministry of Education) College of Physics, Jilin University Changchun, P. R. China Tao Lv College of Physics and Optoelectronic Engineering Shenzhen University Shenzhen, China

Baoning Wang Key Laboratory of Physics and Technology for Advanced Batteries (Ministry of Education) College of Physics, Jilin University Changchun, P. R. China Xiao-Feng Wang Key Laboratory of Physics and Technology for Advanced Batteries (Ministry of Education) College of Physics, Jilin University Changchun, P. R. China Ziwei Wang School of Chemistry and Chemical Engineering Yangzhou University Yangzhou, Jiangsu, China Qin Xu School of Chemistry and Chemical Engineering Yangzhou University Yangzhou, Jiangsu, China State Key Laboratory of Analytical Chemistry for Life Science, School of Chemistry and Chemical Engineering Nanjing University Nanjing, China

xiii

Contibutors

Yuan Yao Central China Normal University Wuhan, Hubei, China Yu-Jia Zeng College of Physics and Optoelectronic Engineering Shenzhen University Shenzhen, China Luhong Zhang College of Physics and Optoelectronic Engineering Shenzhen University Shenzhen, China Wenjie Zhao Key Laboratory of Physics and Technology for Advanced Batteries (Ministry of Education) College of Physics, Jilin University Changchun, P. R. China

Wei-Wei Zhao State Key Laboratory of Analytical Chemistry for Life Science, School of Chemistry and Chemical Engineering Nanjing University Nanjing, China Kui Zhou Institute for Advanced Study Shenzhen University Shenzhen, Guangdong, China Ye Zhou Institute for Advanced Study Shenzhen University Shenzhen, Guangdong, China Yuan-Cheng Zhu State Key Laboratory of Analytical Chemistry for Life Science, School of Chemistry and Chemical Engineering Nanjing University Nanjing, China

1

Introduction to Organic–Inorganic Heterojunction Kui Zhou and Ye Zhou

Heterojunction is a structure formed by combining two different semiconductor materials together, which is an important material foundation for modern electronics and optoelectronics. Combining two or more materials with different original interfaces to form a heterojunction is critical to the design and manufacture of functional devices, so the research of heterojunctions has also become an important subject in the field of materials [1,2]. As early as 1951, A. I. Gubanov [3,4] performed a theoretical analysis of heterojunctions, but it was limited to the difficulty of heterojunction growth technology. In 1960, Anderson [5] made high-quality heterostructures junction for the first time and proposed a more detailed band diagram and theoretical model. In 1963, H. Kroemer and Z. I. Alferov independently proposed the principle of heterojunction lasers [6,7]. In 1969, H. Kroemer and Z. I. Alferov prepared heterojunction lasers that can continuously operate at room temperature [8–10]. This achievement has established the foundation for the development of modern optoelectronics, which was awarded the 2000 Nobel Prize in Physics. In a heterojunction structure, due to the different electrical and optoelectrical parameters of the two semiconductor materials, such as the bandgap width, conductivity type, dielectric constant, refractive index, and absorption coefficient, etc., the behavior of electrons, the interaction of photons and electrons, and some other physical properties are different from those in a single semiconductor material. Therefore, organic–inorganic heterojunctions have attracted more interest and attention. The photoelectric characteristics of heterojunctions can generally be divided into two categories: one is the photocurrent or photovoltaic voltage generated by absorption of photons; the other is the emission of electrons due to current or electric field excitation. Many approaches can be used to create electron–hole pair in the heterojunction, mainly depending on the wavelength of the incident photons. There are usually two important absorption processes that affect the photoelectric properties of a heterojunction: one is the absorption of impurities or interface states to generate free 1

2

Optoelectronic Organic–Inorganic Semiconductor Heterojunctions

electrons or holes; the other is the intrinsic absorption of electrons that transition from the valence band to the conduction band. Due to these processes, the free ­carriers generated at the interface or the transition region of the heterojunction cause the photogenerated current of the heterojunction. In addition to absorbing photons to generate free carriers, photovoltaic voltages can also be generated. The properties of the heterojunction are determined by the composed semiconductor material. Semiconductor materials have been known since the discovery of electrical phenomena in the 18th century. At that time, the materials could be divided into three categories: conductors (ρ ≤ ~10−6 Ω·m), insulators (ρ ≥ 1010 Ω·m), and semiconductors in between. In 1879, the physicist Hall discovered the Hall effect while studying the conductive mechanism of metals. Since then, scientists have used it to study the conductive properties of semiconductor materials. Semiconductor materials are found to have two differently charged carriers that are fewer in number than metals but have higher mobility. Generally, semiconductors can be categorized as inorganic semiconductors and organic semiconductors. In 1910s, inorganic semiconductor materials were used to make cuprous oxide low-power rectifiers and selenium rectifiers. Although a lot of researches have been done on semiconductors, due to the lack of theory in nature, the trial research has made little progress. Until the early 1930s, due to the development of quantum mechanics and the development of the band concept, semiconductor materials provided a solid theoretical basis. The birth of the germanium transistor in 1948 led humans from the era of electron tubes to the semiconductor era. Into the 1960s, the development of silicon integrated circuits was successful, making a leap in the development of the semiconductor industry. While studying silicon-germanium materials, a lot of research was also awakened on other materials, and it was found that compounds formed by group III and group V elements and their multiple compounds are also semiconductors. In the 1970s, various epitaxial growth technologies were developed, and superlattice quantum well and strain layer composites were prepared. Most inorganic semiconductor materials can be single crystals. Single crystals are formed by periodic repeating arrangements of closely spaced atoms. The distance between adjacent atoms is only a few tenths of a nanometer. Therefore, the electrons are different from the free electrons in the vacuum and the electrons in the isolated atom. The electrons in the crystal are in a so-called band state. The energy band is composed of many energy levels. A forbidden band is isolated between the valence band and the conduction band, and electrons are distributed on the energy levels in the energy band. The reason that solids can conduct electricity is the result of the directional movement of electrons under the action of an external electric field. From the perspective of energy band theory, it is the transition of electrons from one energy level to another. When an external electric field is applied, the electrons on the energy band occupied by the electrons can absorb energy from the external electric field and jump to the unoccupied energy level to form a current. This energy band is often called the conduction band. The full band that has been occupied by valence electrons is called the valence band, and the middle is the forbidden band. In semiconductors, both the holes in the valence band and the electrons in the conduction band participate in conduction, which is the biggest difference from metallic conductors.

Introduction to Organic–Inorganic Heterojunction

3

Inorganic materials have the advantages of high light-dark conductivity, high carrier mobility, and long service life as optoelectronic materials, but they have narrow absorption bands, fewer families of materials, and high production costs. Although the appearance of the superlattice makes it possible to improve its optoelectronic performance, it also complicates the device preparation process, increases the device production cost, and makes it difficult to achieve industrial production. Meanwhile, the research of semiconductor materials started with inorganic materials, but due to its own limitations, people soon thought of organic materials naturally, because there are many more types of organic materials than inorganic materials, and inorganic materials with semiconductor properties must be found from them. With the gradual deepening of human research on material science and engineering, people have synthesized a batch of organic materials with semiconductor characteristics and are trying to apply them to the field of traditional semiconductor devices. Compared with inorganic semiconductor materials, organic semiconductors can be processed by the solution method. Therefore, many non-­traditional device processing technologies such as screen printing, inkjet printing, micro-contact printing, and stamps can be used for device fabrication, making it possible to fabricate large area semiconductor devices and lower the cost. Organic semiconductors include molecular crystals, organic complexes, and polymers. Molecular crystal semiconductors have strong covalent bonds inside them, but due to the weak van der Waals forces, the molecules interact with each other and the distance between the molecules is large, which is not conducive to electronic exchange between molecules. Organic molecular crystal semiconductor bandgap Eg = Ic − Ac, where Ic is the ionization energy of the electrons in the crystal and Ac is electron affinity. Molecular crystals are easy to purify (purification methods include recrystallization, column chromatography, and meteorological transmission, etc.), and small molecules can simply obtain single crystal thin films, and this easily available ordering is very beneficial for charge transport. Organic complexes are a combination of a compound with a high electron affinity (electron acceptor) and a compound with a low ionization energy (electron donor), so it is also called charge transfer complex (CT complex) or donor–acceptor complex (DA complex). The charge transfer between the molecules of the material can greatly increase the conductivity of the material. The most common example of semiconducting polymer is a conjugated polymer. A conjugated polymer refers to a repeating unit of a polymer chain composed of an atomic combination polymer having π bond and sp2 hybridization. Some conjugated polymers have semiconductor and even metal properties. They have provided high-performance single-component devices (such as polymer thin-film transistors (TFTs)), and also multifunctional devices with other materials, such as bulk heterojunction (BHJ) solar cells and small molecule/polymer-based TFTs. Recent advances in polymer design, synthesis, and processing enabled remarkable progress in polymer-based device performance. As we know, organic semiconductors and inorganic semiconductors differ greatly in their properties. Generally, organic semiconductors are broadband materials with a bandgap of 2~3 eV, while the bandgap of inorganic semiconductors such as Ge, Si, and GaAs are only 0.66 eV, 1.12 eV, and 1.42 eV at room temperature. At present, most organic semiconductors in air

4

Optoelectronic Organic–Inorganic Semiconductor Heterojunctions

condition exhibit hole transport properties, that is, hole accumulation occurs under negative gate bias; when the gate bias is positive, electron accumulation is rarely found. Very weak van der Waals forces in organic materials are intermolecular interaction forces, which makes organic materials have low dielectric constant and electron mobility. However, scientists can prepare a variety of materials with different optoelectronic properties by changing the positions and types of substituents in organic compounds, which enriches the types of semiconductor optoelectronic materials. At present, since organic semiconductor materials have adjustable energy bands and a wide variety, the great interest are low-cost solution-processed thin films on various robust or flexible substrates in large area scale, which makes them have broad application prospects. However, due to its poor photoelectric performance, material stability, and poor wear resistance, it seems difficult to really commercialize on a large scale. Despite the demonstration of promising prototypes of inorganic or organic semiconductors, it still remains great challenge in their development and optimization for high-performance device. Particularly, it is becoming more and more difficult to integrate desired properties to individual materials to satisfy the increasing demands of multifunctionality for fundamental studies as well as device designs and optimization. Considering the advantages and disadvantages of the two materials, the use of organic/inorganic heterostructures can take the advantages of the two parts which can be fully utilized to improve the photoelectric performance of the materials. With suitable synthesis technology, simplifying the production process may eventually lead to large-scale industrial production. Besides, some new electrical and optoelectrical properties are expected in pre-designed organic–inorganic heterojunctions due to interface effects including energy-level alignment, which is crucial for potential in optoelectrical application. Ideally, the relative position of the band edges is determined by only the Fermi levels EF and the electron affinities χ. However, it has to be corrected in the real cases due to the following interface effects: (1) energy-level shift resulted from dipoles at the organic/inorganic interface by chemical reactions [11]; (2) band bending caused by pinning of the Fermi level at surface or interface states [12]; (3) the type and level of doping of the inorganic semiconductor that have a strong influence on the band alignment. The detail of the energy alignment will be discussed in Chapter 2. In order to get good organic–inorganic heterojunction interface, the advanced surface deposition techniques are required. There are many growth methods for semiconductor heterojunctions. The methods commonly used in the early days are alloy method, sputtering method, vacuum evaporation method, solution growth method, and chemical vapor deposition method; but it is difficult to obtain highquality heterojunctions. In recent years, three methods have been developed, which are relatively fast and easy to obtain good-quality heterojunctions. Molecular beam epitaxy (MBE) refers to epitaxial growth using atomic or molecular beams as a transport source under ultra-high vacuum conditions. During the growth, there is almost no collision between the molecules. The MBE method has the following characteristics: (1) The source and the substrate can be heated and controlled to

Introduction to Organic–Inorganic Heterojunction

5

make the growth at a low temperature; (2) The growth rate is low and easy to control, which is beneficial to the growth of multilayer heterojunctions; (3) During the growth process, the middle surface is in a vacuum, and the equipment can be used for in situ observation and analysis. However, this method requires high equipment and high production costs, and requires a large amount of experimental funds. Chemical vapor deposition is a method for crystal growth from the gas phase using chemical reactions. Among them, metal organic chemical vapor deposition (MOCVD) is developed by the use of metal organic compounds and trimethylaluminum as the source of the three group elements Ga and Al (TMGa and TMAl), and potassium cyanide as the source of the five group elements. An epitaxial layer is formed on the substrate by pyrolysis at 600–750°C. The reason why MOCVD is valued is that it has the following characteristics: (1) Each component participates in the growth and doping in the form of a gas, so the gas flow can be controlled to control the growth speed and the final control of the composition and thickness of the epitaxial layer; (2) The epitaxial layer is grown by thermal decomposition. It is a single-temperature region growth, easy to control with simple production equipment, and conducive to large-scale production; (3) The metal organic source is approximately proportional to the growth rate of the epitaxial layer, so the growth rate of the epitaxial layer can be adjusted by controlling its flow rate. The liquid phase epitaxy method is to place a single crystal substrate in a saturated or supersaturated solution to grow a single crystal layer on the single crystal substrate consistent with the substrate orientation. It is structurally the same as the original single crystal, but with a different material composition, forming a heterojunction. Its main advantages are as follows: (1) Liquid phase growth has a higher growth rate; (2) Growth in a saturated or supersaturated solution requires simple equipment; (3) High crystal purity, low epitaxial layer dislocation density, and good crystal integrity; (4) Wide choice of dopants for liquid phase growth; (5) Easy operation and straightforward industrial production. Most importantly, atomic layer deposition (ALD) is a universal deposition method in inorganic semiconductor industry, which enables the control of thin film down to the atomic level through a gas-phase chemical reaction, since the ALD method can deposit various materials on almost any substrate. On the other hand, as its organic counterpart, molecular layer deposition (MLD), has opened up promising avenues for the fabrication of pure polymeric thin films or inorganic–organic hybrid thin films, which has many advantages such as lower deposition temperatures, tunable thermal stability, and improved mechanical properties [13–15]. However, the application of MLD is limited by the instability of the most organic precursors and the resulting polymeric layers. For this reason, the deposition temperatures are usually low and the corresponding temperature window may be either very narrow or nonexistent. The detail of MLD of organic–inorganic hybrid materials method will be discussed in Chapter 3. As the molecular scale organic–inorganic heterojunction is obtained, the atomicscale techniques, including scanning tunneling microscopy and scanning tunneling spectroscopy (STM and STS), are usually employed to characterize the atomic morphology and electronic properties of the heterojunction, respectively [16]. The STM

6

Optoelectronic Organic–Inorganic Semiconductor Heterojunctions

results can tell us the structure information of the thin film such as structure geometry, crystal orientation, and even atomic defects in the film. The STS results can provide us the electrical information under different bias, for example, current-­ voltage and differential tunneling conductance spectra and STS mapping, etc. This information can assist to analyze the optoelectronic performance based on the organic–inorganic heterojunction. The detail of scanning tunneling microscope and spectroscope on organic–inorganic material heterojunction will be discussed in Chapter 4. As high integration is required in the development of electronics and optoelectronics, the research of low-dimensional material synthesis and micro–nano processing technology is the mainstream trend and research focus. The low-dimensional materials include not only 0D nanodots and quantum dots, 1D nanowires and 2D materials but also low-dimensional organic semiconductors prepared by different processes such as pentacene, copper phthalate (CuPc), Per-3,4,9,10-tetrahydroacid dioxin (PTCDA) and dioctylbenzopyrenebenzobenzophenone (Cg-BTBT), fullerene (C60), rubrene (Rubrene), etc. These low-dimensional materials come from a wide range of sources and have excellent electrical, optical, mechanical, and thermal properties. Moreover, in terms of forming heterojunction with organic semiconductors, the atomically flat surface of low-dimensional materials not only provides an ideal interface for high-efficiency charge separation, transfer, and transport but also act as a template for the epitaxial growth of organic semiconductors [1]. As a consequence, low-dimensional hybrid heterojunction generally exhibits enhanced electrical and optical performances compared to heterostructures consisting bulk phase of organic or inorganic materials. The 2D/organic heterojunctions has been reviewed recently [17]. Thus, the 0D nanodots and 1D nanowires heterojunctions will be discussed in Chapter 5 and Chapter 6, respxectively. Furthermore, an individual molecule has been theoretically predicted as an active part in electronic or optoelectronic devices in 1974 [18]. The first conductance junction based on single molecule was fabricated experimentally in 1997 [19]. Molecular junctions are now extensively studied in both nanoscale devices and fundamental physical properties of molecule materials [20]. Moreover, advanced fabrication process combined with optical technologies has enabled optical experiments on currentcarrying molecular junctions [21,22]. This progress connects molecular electronics with optical spectroscopy together, which has opened up an avenue of molecular optoelectronics [23]. Especially, the scanning tunneling microscope-induced lightemission technology can use the tunneling current of the STM as an atomic-scale source for induction of light emission from a single molecule. Thus, it has enabled the investigation of single-molecule properties at subnanometer spatial resolution. The detail of Electroluminescence of Organic Molecular Junction in Scanning Tunneling Microscope will be introduced in Chapter 7. Based on the basic introduction above, we present recent advances of organic– inorganic semiconductor heterojunction in optoelectronic applications, for example, diodes, solar cells, light-emitting diodes, nanogenerators, resistive memories, photocatalyst, transistors, and wearable sensors in the following chapters of this book (Figure 1.1).

Introduction to Organic–Inorganic Heterojunction

7

FIGURE 1.1  Organic–inorganic heterojunction and related electronic and optoelectronic applications.

REFERENCES 1. Y. L. Huang, Y. J. Zheng, Z. Song, D. Chi, A. T. S. Wee, S. Y. Quek, Chemical Society Reviews 2018, 47, 3241–3264. 2. Y. Liu, Y. Huang, X. Duan, Nature 2019, 567, 323–333. 3. A. Gubanov, Zhurnal Eksperimental'noi i Teoreticheskoi Fiziki Zh. Eksper. Teor. Fiz. 1951, 21, 721–730. 4. J. T. Calow, P. J. Deasley, S. J. T. Owen, P. W. Webb, Journal of Materials Science 1967, 2, 88–96. 5. R. L. Anderson, Experiments on Ge-GaAs Heterojunctions. In Electronic Structure of Semiconductor Heterojunctions, Margaritondo, G., Ed. Springer Netherlands: Dordrecht, 1988; pp 35–48. 6. H. Kroemer, Proceedings of the IEEE 1982, 70, 13–25. 7. Z. I. Alferov, Reviews of Modern Physics 2001, 73, 767-782. 8. Z. I. Alferov, V. Andreev, D. Garbuzov, Y. V. Zhilyaev, E. Morozov, E. Portnoi, V. Trofim, Soviet Physics Semiconductors 1971, 4, 1573–1575. 9. I. Hayashi, M. Panish, P. Foy, IEEE Journal of Quantum Electronics 1969, 5, 211–212. 10. H. Kressel, H. Nelson, RCA Review 1969, 30, 106–113. 11. H. Ishii, K. Sugiyama, E. Ito, K. Seki, Advanced Materials 1999, 11, 605–625. 12. S. M. Sze, K. K. Ng, Physics of Semiconductor Devices. John Wiley & Sons: Hoboken, NJ, 2006. 13. X. Li, A. Lushington, J. Liu, R. Li, X. Sun, Chemical Communications 2014, 50, 9757–9760.

8

Optoelectronic Organic–Inorganic Semiconductor Heterojunctions

14. X. Li, A. Lushington, Q. Sun, W. Xiao, J. Liu, B. Wang, Y. Ye, K. Nie, Y. Hu, Q. Xiao, R. Li, J. Guo, T.-K. Sham, X. Sun, Nano Letters 2016, 16, 3545–3549. 15. Y. Zhao, L. V. Goncharova, Q. Sun, X. Li, A. Lushington, B. Wang, R. Li, F. Dai, M. Cai, X. Sun, Small Methods 2018, 2, 1700417. 16. X. Liu, Z. Wei, I. Balla, A. J. Mannix, N. P. Guisinger, E. Luijten, M. C. Hersam, Science Advances 2017, 3, e1602356. 17. J. Sun, Y. Choi, Y. J. Choi, S. Kim, J.-H. Park, S. Lee, J. H. Cho, Advanced Materials 2019, 31, 1803831. 18. A. Aviram, M. A. Ratner, Chemical Physics Letters 1974, 29, 277–283. 19. M. A. Reed, C. Zhou, C. J. Muller, T. P. Burgin, J. M. Tour, Science 1997, 278, 252. 20. S. V. Aradhya, L. Venkataraman, Nature Nanotechnology 2013, 8, 399–410. 21. T. Shamai, Y. Selzer, Chemical Society Reviews 2011, 40, 2293–2305. 22. R. Zhang, Y. Zhang, Z. C. Dong, S. Jiang, C. Zhang, L. G. Chen, L. Zhang, Y. Liao, J. Aizpurua, Y. Luo, J. L. Yang, J. G. Hou, Nature 2013, 498, 82–86. 23. M. Galperin, Chemical Society Reviews 2017, 46, 4000–4019.

2

Energy-Level Alignment at Organic–Inorganic Heterojunctions Sylke Blumstengel and Norbert Koch

CONTENTS 2.1 Introduction........................................................................................................ 9 2.2 Interface Formation between Organic and Inorganic Semiconductors: 6P on ZnO............................................................ 11 2.3 Work Function Tuning of ZnO via Dipole Bearing Self-assembled Monolayers............................................................................. 14 2.4 Work Function Tuning with Electron Donor and Acceptor Molecules........... 16 2.5 Fingerprint of Ground-State Charge Transfer in the Optical Spectra of ZnO-Acceptor Interfaces................................................... 19 2.6 Organic–Inorganic Semiconductor pn-Junction.............................................. 23 2.7 Energy-Level Tuned Organic–Inorganic Heterojunctions for Light-Emitting Applications............................................ 25 Acknowledgements................................................................................................... 30 References................................................................................................................. 30

2.1 INTRODUCTION The energy-level alignment between inorganic (ISC) and organic (OSC) semiconductors determines the direction and the efficiency of charge transfer and thus the functionality of the heterointerface in an optoelectronic device. For light-emitting applications, for example, a type-I interface (Figure 2.1) would be most suitable. In such a configuration, excitation energy could be transferred not only via excitonic coupling (dipole–dipole interaction) but also by simultaneous injection of electrons and holes into the lower bandgap material (Itskos et al. 2009; Bianchi et al. 2014; Schlesinger et al. 2015). Light-to-electrical energy conversion relies, on the other hand, on efficient exciton dissociation at heterojunctions; therefore, type-II i­ nterfaces with energy offsets larger than the exciton binding energy are required (Figure 2.1) (Oosterhout et al. 2011; Baeten et al. 2011; Eyer et al. 2017; Itskos et al. 2013). Another example is the use of the ISC as carrier injecting or extracting contact in organic electronic devices. Very widespread in applications is ZnO; however, its moderate work function (φ) in the range of about 3.6–4.5 eV often leads to a situa­ tion where the Fermi level (EF), which is close to the conduction band (CB) in n-type ZnO, is located within the energy gap between the highest occupied molecular orbital 9

10

Optoelectronic Organic-Inorganic Semiconductor Heterojunctions

FIGURE 2.1  Schematic depiction of possible energy-level alignments between frontier molecular orbitals and semiconductor band edges at organic–inorganic heterojunctions: (a) type-I and (b) type-II. At type-I interfaces, electrons and holes are captured in the lower bandgap material while type-II interfaces facilitate electron–hole separation. VB: valance band, CB: conduction band.

(HOMO) level and the lowest unoccupied molecular orbital (LUMO) level of the OSC. This results in significant barriers for hole or electron injection (Bhosle et al. 2007; Kim et al. 2013; Bernède et al. 2008). These examples already illustrate the fundamental need of methods for tailoring the relative positions of the occupied and unoccupied energy levels at junctions of ISCs and OSCs to achieve the desired function and to design hybrid optoelectronic devices with indeed superior performance. Due to the rich physics and chemistry involved in interface formation between OSC and ISC, for example, redistribution of charges, interface dipole formation, or chemical bonding, it is virtually impossible to predict the energy offset for a given material pair on the basis of tabulated values of ionization energy (IE) and electron affinity (EA). On the other hand, once the mechanisms responsible for level alignment are understood, this knowledge can be employed to engineer the level offsets at organic–inorganic semiconductor interfaces. This chapter is organized as follows: To provide insight into the processes that govern the energy-level alignment at organic–inorganic semiconductor heterojunctions, a model junction (p-sexiphenyl and ZnO) is presented in Section 2.2 where, in particular, the influence of the termination of the inorganic semiconductor surface and the molecular film structure is discussed. With this basic understanding of the electronic structure at pristine ISC–OSC interfaces, we present methods to tune the interface energetics in order to control charge injection/transfer and achieve specific functions. Adjustment of the interface energetics is possible by the introduction of a suitable molecular interlayer between ISC and OSC. This concept was originally put forward in order to tune the Schottky barrier height of ISC- and OSC-metal junctions and later transferred to OSC-ITO junctions (Campbell et al. 1996; Vilan et al. 2000, 2010; de Boer et al. 2005; Koch et al. 2005a; Koh et al. 2006; Bröker et al. 2008; Alloway et al. 2009; Rangger et al. 2009; Niederhausen et al. 2011; Hotchkiss et al. 2012; Asyuda et al. 2020). We discuss two fundamental approaches: (i) Chemisorption of molecules with an electric dipole moment perpendicular to the interface on the inorganic semiconductor surface (Section 2.3) (Hotchkiss et al. 2011; Lange et al. 2014; Kedem et al. 2014; Timpel et al. 2014, 2015). The approach is exemplified for

Energy-Level Alignment at Organic–Inorganic Heterojunctions

11

various ZnO surfaces functionalized with self-assembled monolayers (SAMs) with varying dipole moment. Changing the composition of the dipole layer allows a finetuning of the work function in a moderate energy range (≤1.5 eV); (ii) Introduction of a monolayer of strong molecular acceptors or donors (Schlesinger et al. 2013; Schlesinger et al. 2015, 2019a; Schultz et al. 2016). With this approach, a wider tuning of the work function of wide bandgap semiconductors (ZnO, GaN) is obtained spanning a range from 2.3 eV to 6.5 eV (Section 2.4). Substantial charge redistribution at the heterojunction is responsible for the strong modulation of φ. This also affects strongly the optical properties of the semiconductor surface, which in turn opens up an alternative way to study the electronic structure at organic–inorganic interfaces (Meisel et al. 2018). In Section 2.5, it is shown that the analysis of the evolution of the optical spectra of the ISC surface upon deposition of acceptor molecules provides an estimate of the magnitude of the electrostatic potential change. The effect of p-doping of an OSC on the energy-level alignment at the interface to the natively n-type ISC (ZnO) is discussed in Section 2.6 (Futscher et al. 2019), that is, a prototypical pn-junction. Finally, Section 2.7 demonstrates that the energy-level tuning via a molecular interlayer provides indeed the desired control over the charge transfer processes at the junction and its functionality (Schlesinger et al. 2015). In this example, a strong increase of the luminescence yield is achieved rendering such well-tailored heterojunctions suitable for light-emitting applications.

2.2 INTERFACE FORMATION BETWEEN ORGANIC AND INORGANIC SEMICONDUCTORS: 6P ON ZNO The simplest model to predict how the energy levels of two semiconductors line up when brought into contact is based on an assumed vacuum level (E0, where the electrostatic potential is set to zero) alignment across the interface. Accordingly, the electrostatic potential across the interface is constant and this implies that no charge density rearrangement occurs upon contact, that is, the limit of weak van der Waals interactions or physisorption. This model further neglects eventual doping of the semiconductors, as well as surface and interface states within the otherwise empty bandgaps. In that simple case, knowing the IE and EA of the two materials is sufficient to predict the offset between unoccupied (Δunocc) and occupied (Δocc) energy levels, as shown in Figure 2.2. While this situation is, in fact, often encountered at purely OSC heterojunctions, it is scarcely found when an ISC is involved. The main reason is that for ISCs there is a finite density of electrons spilling out into vacuum at the free surface, forming a net surface dipole that increases the sample work function, IEISC, and EAISC, by the same amount. This is in full analogy to the surface dipole of metals (Smoluchowski 1941), which is, however, larger in magnitude compared to ISCs. This electron density tailing out of the free surface becomes “pushedback” by the adsorption of molecules, which effectively reduces the surface dipole and thus results in a reduction of electrostatic potential at the interface by ΔE0. This mechanism has been first understood for metal–OSC interfaces (Koch et al. 2003), but it was also found for physisorptive ISC–OSC junctions (Greiner et al. 2012; Winkler et al. 2013; Schlesinger et al. 2019b). In literature, this phenomenon is often termed “interface dipole”. However, the use of this term should be avoided when

12

Optoelectronic Organic-Inorganic Semiconductor Heterojunctions

FIGURE 2.2  Schematic energy-level diagrams of an ISC–-OSC heterojunction with (a) vacuum level alignment and (b) with an abrupt shift of the vacuum level (E0) across the inter face by ΔE0. The latter changes the energy offset between the unoccupied levels (Δunocc, Δ'unocc) and occupied levels (Δocc, Δ'occ) at the interface, compared to vacuum level alignment. VBM and CBM are the valence band maximum and conduction band minimum of the ISC, respectively.

describing the interfacial “push-back”, because no dipole is formed upon contact of the materials, but instead an already existing one (the surface dipole) is modified, as experimentally observed as a change of sample φ. The consequence of this on the level alignment is a rigid shift of the HOMO and LUMO levels of the OSC compared to the ISC by ΔE0, and thus different level offsets Δ'unocc and Δ'occ, as shown in Figure 2.2. However, many actual ISC–OSC junctions can feature much more complexity. For instance, φ of the ISC surface depends on crystallographic orientation of the surface [just like φ of metals (Smoluchowski 1941)] and surface termination. Furthermore, IEOSC and EAOSC are not intrinsic material properties, but they depend on the relative orientation of the molecular assembly with respect to an interface (Duhm et al. 2008), with variations reaching up to 1 eV or more. The impact of both material-specific property variations on the interfacial energy levels is exemplified in the following for the physisorptive interfaces between para-sexiphenyl (6P) and three different ZnO surfaces (Schlesinger et al. 2019b). The bare surface work function values are 3.6 eV for ZnO(0001), 3.9 eV for ZnO(101; 0), and 4.3 eV for ZnO(0001; 0 ), and all ZnO crystals are n-doped. The rodlike molecule 6P (chemical structure shown in the inset of Figure 2.3) exhibits a growth mode on all three ZnO surfaces that is characterized by a flat-lying (L) first layer and essentially upright standing (S) second and higher layers. This change in relative molecular orientation with respect to the substrate plane as function of layer is due to the fine balance of substrate-monolayer (dominant for the monolayer) and intermolecular interactions (dominant in multilayers). As noted above, the change in molecular orientation goes hand in hand with a change of the IE and EA of 6P by

13

Energy-Level Alignment at Organic–Inorganic Heterojunctions

6P

FIGURE 2.3  Schematic energy-level diagrams for 6P on (a) ZnO(0001; 0 ), (b) ZnO(101; 0 ), and (c) ZnO(0001). ZnO levels are for bare surfaces prior to adsorption. Adsorption-induced changes of ZnO levels (due to surface band bending changes) are indicated by arrows. The first 6P layer on ZnO is lying (“L”) face-on, which leads to energy-level pinning on ZnO(101; 0 ) and ZnO(0001) raising φ to 4.0 eV. Subsequent layers of 6P are standing (“S”) and vacuum aligned with the lying layer underneath. For the “L” layer the diagrams differentiate between the energy-level positions due to the different orientations alone (ΔIE/EA – thin lines), and including the narrowing of the energy gap (boxes ΔIE/EA+ΔEg – thick lines); dashed boxes (w/o Δφ) indicate the LUMO states position before interfacial charge transfer due to EF-pinning. Adapted from (Schlesinger et al. 2019b).

ΔIE/EA ≈ 0.8 eV, which by itself would already have a notable effect on the interface energetics. In addition, the monolayer experiences two more effects: (i) the proximity to the ZnO (with a dielectric constant of ca. 8, compared to that of 6P of ca. 3) lowers the energy gap by ca. 0.3 eV due to increased dielectric screening compared to 6P molecules in the bulk; (ii) the inter-ring twist angle of 6P in the bulk is smaller for the adsorbed monolayer and that additionally lowers the energy gap by up to 0.3 eV (Koch et al. 2005b). In combination, the energy gap of the lying 6P monolayer on ZnO is reduced by up to ΔEg ≈ 0.6 eV compared to the bulk with an energy gap of ca. 3.5 eV (Hwang, Wan, and Kahn 2009). Let us first consider the highest φ ZnO surface, that is, ZnO(0001; 0 ) with 4.3 eV. Deposition of 6P reduces φ by 0.2 eV and the Fermi level (EF) of the n-doped ZnO is located well within the energy gap of 6P for both lying mono- and standing multilayer, as shown in Figure 2.3a. This situation thus corresponds to the scheme shown in Figure 2.2b, that is, an interface without charge transfer and only the “push-back” effect that lowers the electrostatic potential (in experiments observed as work function change). Turning toward the ZnO(101; 0) surface with an initial φ of 3.9 eV, we observe that 6P deposition slightly increases φ by 0.1 eV, and an even larger work function increase by 0.4 eV is found for ZnO(0001) with an initial φ of 3.6 eV. Since at least a small φ lowering is expected due to the push-back effect, the opposite change of the work function thus indicates that another mechanism must be at play for these two interfaces. The initial φ values of ZnO(101; 0) and ZnO(0001) are sufficiently low so that EF would come to lie within the LUMO level manifold of the lying 6P layer due to the EA increase brought about by the gap narrowing effects of screening and reduced inter-ring twist angle (see preceding paragraph). To reach

14

Optoelectronic Organic-Inorganic Semiconductor Heterojunctions

electronic equilibrium, electrons are thus transferred from the ZnO to 6P molecules. This results in an increase of φ, as seen in experiment, and concomitantly shifts the energy levels of all 6P molecules upward, as schematically shown in Figure 2.3. This interfacial charge transfer at a physisorptive interface, conventionally termed Fermilevel pinning (Yang et al. 2017; Mao et al. 2011), is driven by electronic equilibrium only so that no unoccupied levels come to lie below EF. The above examples demonstrate how the different simultaneously acting mechanisms of push-back, Fermi-level pinning, dielectric screening, molecular orientation change, and molecular conformation dependent energy gap narrowing are responsible for essentially identical level alignment between ZnO and multilayer 6P, despite varying starting conditions. It is rather difficult to access the information on the presence and properties of such molecularly thin interlayers, like the lying 6P layer here. If these details of the ISC–OSC junction go unnoticed, correlations between functional behavior of devices and assumed interfacial energy levels may lead to false conclusions for further development.

2.3 WORK FUNCTION TUNING OF ZNO VIA DIPOLE BEARING SELF-ASSEMBLED MONOLAYERS The most straightforward way to tune the energy-level alignment between ISC and OSC is the introduction of a monolayer of dipole bearing molecules between the two materials. If these molecules are aligned, the resulting dipole layer just resembles a parallel plate capacitor and the potential change ∆φ perpendicular to the surface can be calculated by the Helmholtz equation

 

e    d e  n    (2.1) r   0 r   0

where e is the elementary charge, σ the area charge density, d the thickness of the dipole layer, εr and ε0 are the relative and vacuum dielectric permittivity, respectively. n is the area density of the dipoles and μ⊥ the dipole moment component perpendicular to the interface. It comprises contributions of the molecular dipole moment as well as the bonding dipole. Well-ordered dipole layers are achieved by the wet-chemical deposition of SAMs consisting of rod-like molecules bearing on one side an anchoring group capable of binding to the oxide surface and on the other side an either electron withdrawing or electron donating tail group. Both groups are connected by a molecular spacer, which, together with the anchoring group, determines the packing density and the structure of the monolayer. The molecules discussed in the following possess as a molecular spacer a polarizable benzene ring and as anchor a phosphonic acid group. The phosphonate head group is known to have strong affinity to ZnO so that highly dense molecular films can be prepared (Ostapenko et al. 2016). A XPS analysis of the O 1s core level shows that the molecules adsorb on ZnO via a mixture of bidentate and tridentate binding configurations (Figure 2.4a) independent of the tail group (Timpel et al. 2015). The tuning of the work function of ZnO is exemplary demonstrated with the three phenyl

Energy-Level Alignment at Organic–Inorganic Heterojunctions

15

FIGURE 2.4  (a) Schematic illustration of the phosphonic acid binding modes on ZnO. (b) Work function modification ∆φ as a function of the Hammett parameter of the molecules depicted in the inset. The direction of the molecular dipole moment is indicated as well. The data points refer to sol–gel ZnO (red), ALD ZnO (green), and MBE grown ZnO(0001) (blue). (c) Work function tuning using SAMs of a mixture of PO3-OCH3 and PO3-CN. The difference in the work function between the mixed SAM and the pure PO3-OCH3 SAM is plotted as a function of the partial concentration cPO3 − CN of PO3-CN in the solution used to prepare the SAM. The total concentration of the two molecules is kept constant. (b) and (c) Adapted from (Kedem et al. 2014).

phosphonates: 4-methoxyphenyl phosphonic acid (PO3-OCH3), phenyl phosphonic acid (Phen-PO3), and 4-cyanophenyl phosphonic acid (PO3-CN), differing in the tail groups and thus in the molecular dipole moment (Kedem et al. 2014). The chemical structure of the molecules is shown in the inset of Figure 2.4b. To investigate if the ZnO surface termination, the morphology, and the intrinsic doping level affect the work function tuning, ZnO films are prepared by atomic layer deposition (ALD), sol–gel deposition, and molecular beam epitaxy (MBE). While ALD and sol–gel deposition results in polycrystalline films, MBE yields single crystalline O-terminated ZnO(0001). Changes in the work function have been recorded by combined Kelvin Probe and UPS measurements. The results are summarized in Figure 2.4b where the work function changes ∆φ as a function of the Hammett substituent parameter of the molecules are depicted. This parameter describes the tendency of the functional group (OCH3, H and CN in this case) to donate or withdraw electrons from the benzene ring and is thus a measure of the molecular dipole moment. As seen in Figure 2.4b, a linear correlation of the work function with the Hammett parameter is found for all ZnO morphologies, regardless of the totally achieved range, indicating that the effect of the molecules on the various types of ZnO films is essentially the same. The electron-donating methoxy tail group results in a positive dipole, pointing out from the surface, which decreases the work function, while the electron-withdrawing cyano tail group introduces a negative dipole, which increases the work function. The shift with the unsubstituted PO3-Phen termination allows separation of the effect of the substituents alone from the effect of the dipole due to the bonding of the molecule to the surface, which changes the ZnO termination from Zn-OH to Zn-O-P, as well as any additional dipole caused by the unsubstituted molecule. A maximum work function span ∆φ of ~0.8…1.1 eV between the opposite dipoles is obtained. The small differences in the totally achieved ∆φ on the different ZnO surface terminations and

16

Optoelectronic Organic-Inorganic Semiconductor Heterojunctions

morphologies can be related to a slight modification of dipole density and/or angle. To address the question, if the dipole layer is solely responsible for the work function modification or if a changed surface band bending within ZnO contributes, XPS measurements were performed (Kedem et al. 2014; Timpel et al. 2015). A shift in the Zn 2p or 3s peak positions due to assembly of the molecular layer would indicate a change in band bending. However, the Zn 2p and 3s peak positions stayed constant at the values of the bare ZnO (~1022 eV and ~140 eV, respectively). The major effect is thus extrinsic to ZnO and due to the molecular layer’s dipole. By mixing of PO3-OCH3 and PO3-CN which possess opposite dipoles, the work function can be varied in a controllable manner according to the concentration ratio of the two molecules in the solution used to form the SAM as shown in Figure 2.4c. The work function is found to correlate linearly with the partial concentration of PO3-CN in the solution. The advantage of such an approach is that only two types of molecules are required to span the entire modification range. The approach thus allows a continuous modification of the energy-level alignment at the heterointerface between ISC and OSC. Introduction of dipolar SAMs presents a viable route to tune the work function of technologically relevant ZnO surface. The expansion of the functionality due to SAM modification has been demonstrated, for example, in planar hybrid photovoltaic diodes with the organic donor, N,N′-di(1-naphthyl)-N,N′-diphenyl-(1,1′biphenyl)-4,4′-diamine (α-NPD) and ZnO as acceptor (Piersimoni et al. 2015). The open circuit voltage VOC at organic–inorganic heterojunctions is correlated to the hybrid energy gap between CBM of the ISC and the HOMO of the OSC. By introduction of phenyl phosphonates with different tail groups as well as mixtures thereof, the hybrid energy gap at the α-NPD/ZnO heterojunction was varied by 300 meV. It was shown, that the VOC of the devices changes by the same amount. Applying a forward bias to the devices, near-infrared electroluminescence is emitted stemming from the recombination of electrons in the ZnO CB with holes in HOMO of α-NPDs at the heterojunction. Similar to the VOC, the peak position of the electroluminescence is also tunable by the introduction of the SAMs. Another example for the expansion of the functionality is the application of SAM-modified ZnO as injecting electrodes in optoelectronic devices (Lange et al. 2014). In that work, it is shown that ZnO can serve both as electron- and hole-injecting contact, and furthermore that the injection properties can be continuously altered from being strongly injection limited to Ohmic. Consequently, unipolar currents in P3HT and phenyl-C71-butyric acid methyl ester (PCBM)-based diodes could be tuned by several orders of magnitude just by controlling the ZnO work function with an appropriate SAM.

2.4 WORK FUNCTION TUNING WITH ELECTRON DONOR AND ACCEPTOR MOLECULES Another approach to tune the work function of ISC surfaces, in order to modify the level alignment with a subsequently deposited OSC as discussed in Section 2.3, builds on a concept developed earlier to modify φ of metal electrodes for reducing charge injection barriers into OSCs. Molecules with strong electron donor and acceptor character chemisorb on metal surfaces involving pronounced charge transfer.

Energy-Level Alignment at Organic–Inorganic Heterojunctions

17

For  instance, the strong molecular acceptor tetrafluoro-tetracyanoquinodimethane (F4TCNQ) binds to the Ag(111) surface via orbital hybridization and bi-directional electron transfer involving numerous orbitals (Romaner et al. 2007). Overall, ca. one electron is transferred from the metal to F4TCNQ, the positive counter-charge is located at the Ag surface. Consequently, each molecule-metal complex has a net dipole moment perpendicular to the interface, which increases φ according to the Helmholtz equation above. A decrease of the metal surface φ can be readily achieved by using a strong molecular donor (Bröker et al. 2008). Such acceptor- or donor-induced interfacial charge transfer also occurs with many ISCs. For instance, Figure 2.5a reveals that huge φ increases can be achieved by depositing the acceptors 1,3,4,5,7,8-hexafluoro-tetracyanonaphthoquinodimethane (F6TCNNQ) and 1,4,5,8,9,11-hexaazatriphenylenehexacarbonitrile (HATCN) onto ZnO and GaN. However, the mechanism leading to Δφ is different from the metal case because the semiconducting nature of the material must be accounted for. First of all, no indications for pronounced chemisorption and orbital hybridization of such acceptors on semiconductor surfaces were found for the cases studied to date, including hydrogen-terminated Si surfaces (Wang et al. 2019), and thus integer charge transfer across the interface is considered (Schöttner et al. 2020) – in analogy to Fermi-level pinning discussed in Section 2.2. Next, it is generally observed that the overall Δφ is in part due to molecule-induced (modified) surface band bending within the ISC, termed ΔφBB and exemplarily shown in Figure 2.5b for the same samples where Δφ is shown in Figure 2.5a. At this point, we recall that ISCs are generally doped and possess a certain gap-state density of states (GDOS) at their surface, and

FIGURE 2.5  Work function change Δφ (a) and change in surface band bending ΔφBB (b) for GaN(0001) and ZnO(0001) upon stepwise deposition of the acceptors HATCN and F6TCNNQ. Valence photoemission spectra for different ZnO surfaces, plotted on (c) linear and (d) logarithmic intensity scale. The onset positions of the VBM are indicated with red markers. The breaks in (d) indicate concatenation of individually measured spectra. On logarithmic intensity scale a continuous (decaying toward EF, set to zero) photoemission intensity in the forbidden energy gap region of ZnO (GDOS) is visible. For ZnO(101; 0 ) and the highly Ga-doped ZnO(0001; 0 ) face, the band tails fall off approximately exponentially (linear slope in logarithmic scale; indicated by dashed lines).

18

Optoelectronic Organic-Inorganic Semiconductor Heterojunctions

both can be the source of electrons (holes) that are transferred to the molecular acceptors (donors). In the following, the case of acceptors deposited on an n-doped ISC is discussed specifically, since this represents the most case studies to date, often with ZnO as ISC. The presence of a GDOS cannot be readily inferred from a photoemission valence spectrum plotted on a linear intensity scale (Figure 2.5c), as the high density of states (DOS) from the valence band dominates and the region between the VBM (indicated by red markers) and EF appears empty. The same spectra plotted on logarithmic intensity scale (Figure 2.5d), however, allow directly observing that the region of the supposedly empty bandgap does feature a finite GDOS, whose energy and intensity distribution depends on details of the sample surface and may extend throughout the gap up to the CB. Here, for the highly doped ZnO(0001; 0 ) and Zn(0001) even part of the CB is filled with electrons. Thus, in addition to electrons from ionized shallow donors from the bulk of the n-doped ISC, also electrons from this GDOS are transferred to the adsorbed molecular acceptors. Both, however, contribute differently to the overall work function change Δφ, as explained below. The overall density of electrons transferred to the molecular acceptor layer is

2eN q   0   EAmm / kBT (2.2) e 1

with e the elementary charge, Nm the area density of acceptors, φ0 the bare ISC work function, EAm the electron affinity of the acceptors, kB the Boltzmann constant, and T the temperature. Thus, electron transfer stops as soon as no more LUMO levels of the acceptor layer are below EF. The overall work function change has then two contributions, due to band bending in within the ISC (ΔφBB) and an interface dipole between the charged acceptors and the charges in the ISC (ΔφID) (Schultz et al. 2016):

   BB  ID (2.3)

with 2



 BB

  ND  q N  N  D GDOS   (2.4) 20N D

and

ID  e

q deff (2.5) 0

where ND is the donor density; NGDOS is the density of surface gap states; ε0 is the vacuum permittivity; ε is the dielectric constant of the inorganic semiconductor; and deff is an effective distance between the acceptor molecules and the surface of the inorganic semiconductor. The conceptual partitioning of Δφ into ΔφBB and ΔφID is summarized in Figure 2.6. For practical ISCs one can state that the contribution of band bending to the overall work function becomes smaller for higher doping level (ND) of the semiconductor,

Energy-Level Alignment at Organic–Inorganic Heterojunctions

19

FIGURE 2.6  Schematic illustration of charge density (ρ) distribution, work function change (Δφ), and resulting energy-level diagrams for molecular acceptors adsorbed on an n-type ISC. Shallow donors in the bulk of the ISC with a binding energy ED below the CB.

FIGURE 2.7  Representation of the achieved range of work function tuning by moelcular acceptors and donors for ZnO, compared to the work function range of the bare ISC.

and it is further reduced by an increasing gap-state density (NGDOS). In the limit of a metal instead of a semiconductor, only ΔφID is responsible for the work function change. Considerations regarding molecular donors on ISCs to reduce the work function are analogous to the above. Overall, it is quite remarkable to which extent φ of ISCs can be tuned by molecular donors and acceptors, as shown in Figure 2.7 for ZnO. This, in turn, would allow for concomitant energy-level adjustment between an ISC and an OSC with a thin (monolayer range) interlayer of acceptors or donors, and tuning by over 1 eV in either direction has already been demonstrated (Schlesinger et al. 2013, 2015).

2.5 FINGERPRINT OF GROUND-STATE CHARGE TRANSFER IN THE OPTICAL SPECTRA OF ZNO-ACCEPTOR INTERFACES As shown in the previous section, the huge work function increase induced at wide bandgap ISC surfaces by adsorption of acceptor molecules is due to charge redistribution. The change in the electrostatic potential is due to the formation of an interface dipole as well as a strong band bending in the ISC. In this section, it is shown that the electric field associated to the band bending (Figure 2.8a) massively changes the optical spectra of the ISC surface (Meisel et al. 2018). This intricate interplay between the electronic and the optical properties has to be, on the one hand, taken

20

Optoelectronic Organic-Inorganic Semiconductor Heterojunctions

FIGURE 2.8  DR spectra of (a) the F6TCNNQ/ZnO(0001) and (b) the F6TCNNQ/Al2O3(0001) interface for increasing thickness of the organic layer. (c) Simulated DR spectra of a F6TCNNQ/ ZnO (black) and a F6TCNNQ/F6TCNNQ2-/ZnO (red) interface using the dielectric functions of the individual components. The simulations are performed using the transfer matrix method using the dielectric functions of the individual components. (b)–(d) Adapted from Meisel et al. (2018).

into account when designing heterojunctions for specific optoelectronic functions. On the other hand, if well understood, it can be employed to reveal the occurrence of charge redistribution at organic–inorganic heterojunctions and to study in situ and in real time the electronic structure evolution. We will show in the following examples that the analysis of the optical spectra can even yield an estimate of the magnitude of the electric field and consequently of the change in the electrostatic potential due to molecular adsorption. Optical spectroscopy thus provides a simple alternative to photoemission spectroscopy. The approach is presented for the example of the just discussed F6TCNNQ-ZnO heterojunction. The changes in the optical spectra are tracked by differential reflectance (DR) spectroscopy during the vacuum deposition of the acceptor molecules on the ZnO surface. The DR signal is defined as

R R  d   R  0  (2.6)  R R 0

where R(0) is the reflection spectrum of the pristine semiconductor surface recorded prior to the deposition of the molecules, while R(d) refers to the reflection spectra of the surface covered by a molecular layer with the thickness d. Since the difference spectrum is measured in situ during the sublimation of the molecules, DR

Energy-Level Alignment at Organic–Inorganic Heterojunctions

21

spectroscopy is highly sensitive and allows the detection of the signatures of even a fraction of a molecular monolayer (Proehl et al. 2005). Figure 2.8b and c show the evolution of the DR spectra while F6TCNNQ layers grow on a ZnO(0001) surface and as reference on an inert Al2O3 substrate. The reference DR spectra feature a broad unstructured peak centered around 2.7 eV which corresponds to the absorption of the molecule. Assuming that the molecular layer thickness is much smaller than the wavelength of light d ≪ λ, the DR spectra are approximately (McIntyre 1971)

R 8d  1  ˆorg  Im  R   1  ˆinorg

  (2.7) 

with ˆorg and ˆinorg being the complex dielectric functions of the organic layer and of the inorganic substrate, respectively. It is apparent from the equation that in spectral regions where ˆinorg does not vary significantly (i.e., on the transparent Al2O3 in the whole considered spectral range), the DR spectra resemble the imaginary part Im ˆorg of the organic adsorbate layer and thus its absorption spectrum. The broad feature is also visible in the DR spectra of the F6TCNNQ/ZnO(0001) interface and can also here be assigned to the F6TCNNQ absorption since ZnO is transparent in this spectral range. The slight blueshift is explained by a different packing of the molecules and a different dielectric environment. There are three additional sharp features in the spectral range between 3.2 and 3.4 eV which are absent in the reference spectra. They occur already at the lowest coverage and do not grow further in intensity upon deposition of organic material. These features cannot be reproduced by simulations of the DR spectra in the frame of a simple three-layer system consisting of a ZnO half space on one side, a F6TCNNQ layer of thickness d with an oscillator at 2.7 eV accounting for the molecules absorption, and vacuum on the other side (Figure 2.8c). Even if a layer of charged F6TCNNQ2− molecules is introduced, which absorb in the spectral range of interest, that is between 3.2 and 3.4 eV, no agreement with the experimental spectra can be obtained (Figure 2.8d). These features must therefore be a sign for a modification of the dielectric function of ZnO as a consequence of the electronic interaction with F6TCNNQ. The photoemission spectroscopy experiments presented in Section 2.4 provide a straightforward explanation. Remember that the band bending at the ZnO surface changes by ΔΦBB ≈ 1 eV upon deposition of F6TCNNQ (Schultz et al. 2016). The Schottky approximation yields a corresponding electric field F(z  =  0)  ≈  3.7  ×  107 V/m at the ZnO surface which drops linearly over the space charge region of a thickness z ≈ 55 nm to zero (Figure 2.8a) assuming a donor density of ND ≈ 3 × 1023 m−3 and a static dielectric constant of 8 which are reasonable values for the used ZnO. F(z = 0) is thus comparable to the ionization field of the ZnO Wannier–Mott exciton which is FI = 2.6 × 107 V/m. FI corresponds to the potential drop of the effective Rydberg (exciton binding energy) over the exciton Bohr radius. The strong electric field present in the space charge region modifies the shape of the excitonic absorption edge of ZnO and causes the characteristic DR features between 3.2 and 3.4 eV. The quantitative analysis of the DR spectra provides an estimate of the magnitude of the field as briefly

22

Optoelectronic Organic-Inorganic Semiconductor Heterojunctions

outlined in the following. Details on the calculations can be found elsewhere (Meisel et al. 2018). Within the effective mass approximation, the ZnO Wannier–Mott excitons near the surface behave formally like hydrogen atoms in an electric field (Figure 2.9a). Calculated Im ˆZnO spectra as a function of the electric field strength are reported in Figure 2.9b. Qualitatively, the effect of the electric field on the absorption line shape can be understood as follows: At small field strengths F ≪ FI, the electric field leads to a widening of the Coulomb potential well causing a slight redshift of the excitonic transition energy (second order Stark effect). As the field strength increases but F  3.55 eV), more obviously so when the difference spectrum with respect to L4P-sp3 on sapphire is constructed. The lifetime shortening of the QW excitons due to the L4P-sp3 overlayer is basically the same in structure (ii) and (iii) (Figure 2.13f) although the donor–acceptor spatial separation is widened by the ca. 0.3 nm thickness of the [RuCp*mes]+ interlayer. Thus, PLE and time-resolved PL data concordantly yield an efficiency of ηCT= 0.65 for the optimized structure (iii). Fully consistent with the increase in L4P-sp3 lifetime and the unchanged FRET, the molecular emission in the optimized hybrid structure (iii) increases by a factor of seven compared to structure (ii) (Figure 2.13d). The yield of photons emitted by the L4P-sp3 layer per electron–hole pair generated in the QW (either by optical or electrical excitation) is η = ηFRET ∙ ηPL, L4p − sp3. The latter quantity is the emission yield of L4P-sp3 in the hybrid structure. As there is still residual exciton quenching at the ZnMgO interface, ηPL, L4p − sp3≈ 0.55 assuming that the intrinsic PL yield of L4P-sp3 approaches unity. Hence, the total luminescence yield of the hybrid structure (iii) is η ≈ 0.35. At room temperature, the role of the [RuCp*mes]+ interlayer is even more crucial. Whereas L4P-sp3 emission from structure (ii) is no longer detectable, the signal remains bright in case of structure (iii). Despite this impressive improvement of radiative emission yield – particularly at room temperature – the present HIOS can further be optimized, since a considerable fraction of excitons is still not used for light emission. This might be traced back to interface states, which are too low in intensity to be directly revealed by photoemission. To further optimize such hybrid structures, work should focus on identifying and avoiding non-radiative side channels.

ACKNOWLEDGEMENTS This work was funded by the Deutsche Forschungsgemeinschaft (DFG) Projektnummer 182087777 - SFB 951.

REFERENCES Alloway, Dana M, Amy L Graham, Xi Yang, Anoma Mudalige, Jr. Ramon Colorado, Vicki H Wysocki, Jeanne E Pemberton, T. Randall Lee, Ronald J Wysocki, and Neal R Armstrong. 2009. “Tuning the Effective Work Function of Gold and Silver Using

Energy-Level Alignment at Organic–Inorganic Heterojunctions

31

ω-Functionalized Alkanethiols: Varying Surface Composition through Dilution and Choice of Terminal Groups.” The Journal of Physical Chemistry C 113 (47): 20328. doi:10.1021/jp909494r. Asyuda, Andika, Michael Gärtner, Xianglong Wan, Ines Burkhart, Torben Saßmannshausen, Andreas Terfort, and Michael Zharnikov. 2020. “Self-Assembled Monolayers with Embedded Dipole Moments for Work Function Engineering of Oxide Substrates.” The Journal of Physical Chemistry C 124 (16): 8775–8785. doi:10.1021/acs.jpcc.0c00482. Baeten, Linny, Bert Conings, Hans-Gerd Boyen, Jan D’Haen, An Hardy, Marc D’Olieslaeger, Jean V. Manca, and Marlies K. Van Bael. 2011. “Towards Efficient Hybrid Solar Cells Based on Fully Polymer Infiltrated ZnO Nanorod Arrays.” Advanced Materials 23 (25): 2802. doi:10.1002/adma.201100414. Bernède, J C, L Cattin, M Morsli, and Y Berredjem. 2008. “Ultra-Thin Metal Layer Passivation of the Transparent Conductive Anode in Organic Solar Cells.” Solar Energy Materials and Solar Cells 92 (11): 1508–1515. doi:10.1016/j.solmat.2008.06.016. Bhosle, V, J T Prater, Fan Yang, D Burk, S R Forrest, and J Narayan. 2007. “Gallium-Doped Zinc Oxide Films as Transparent Electrodes for Organic Solar Cell Applications.” Journal of Applied Physics 102 (2). doi:10.1063/1.2750410. Bianchi, F, S Sadofev, R Schlesinger, B Kobin, S Hecht, N Koch, F Henneberger, and S Blumstengel. 2014. “Cascade Energy Transfer versus Charge Separation in LadderType Oligo(p-Phenylene)/ZnO Hybrid Structures for Light-Emitting Applications.” Applied Physics Letters 105 (23): 233301. doi:10.1063/1.4903517. Blumstengel, S, S Sadofev, C Xu, J. Puls, R L Johnson, H Glowatzki, N Koch, and F Henneberger. 2008. “Electronic Coupling in Organic-Inorganic Semiconductor Hybrid Structures with Type-II Energy Level Alignment.” Physical Review B - Condensed Matter and Materials Physics 77 (8): 085323. doi:10.1103/PhysRevB.77.085323. de Boer, B, A Hadipour, M M Mandoc, T van Woudenbergh, and P W M Blom. 2005. “Tuning of Metal Work Functions with Self-Assembled Monolayers.” Advanced Materials 17 (5): 621. doi:10.1002/adma.200401216. Bröker, B, R P Blum, J Frisch, A Vollmer, O T Hofmann, R Rieger, K Müllen, J P Rabe, E Zojer, and N Koch. 2008. “Gold Work Function Reduction by 2.2 EV with an Air-Stable Molecular Donor Layer.” Applied Physics Letters 93 (24): 243303. doi:10.1063/1.3049616. Campbell, I H, S Rubin, T A Zawodzinski, J D Kress, R L Martin, D L Smith, N N Barashkov, and J P Ferraris. 1996. “Controlling Schottky Energy Barriers in Organic Electronic Devices Using Self-Assembled Monolayers”. Physical Review B - Condensed Matter and Materials Physics 54 (20): 14321. Duhm, Steffen, Georg Heimel, Ingo Salzmann, Hendrik Glowatzki, Robert L Johnson, Antje Vollmer, Jürgen P Rabe, and Norbert Koch. 2008. “Orientation-Dependent Ionization Energies and Interface Dipoles in Ordered Molecular Assemblies.” Nature Materials 7 (4): 326–332. doi:10.1038/nmat2119. Eyer, Moritz, Johannes Frisch, Sergey Sadofev, Norbert Koch, Emil J W List-Kratochvil, and Sylke Blumstengel. 2017. “Role of Hybrid Charge Transfer States in the Charge Generation at ZnMgO/P3HT Heterojunctions.” Journal of Physical Chemistry C 121 (40): 21955–21961. doi:10.1021/acs.jpcc.7b07293. Futscher, Moritz H, Thorsten Schultz, Johannes Frisch, Maryline Ralaiarisoa, Ezzeldin Metwalli, Marco V Nardi, Peter Müller-Buschbaum, and Norbert Koch. 2019. “Electronic Properties of Hybrid Organic/Inorganic Semiconductor Pn-Junctions.” Journal of Physics Condensed Matter 31 (6): 064002. doi:10.1088/1361-648X/aaf310. Ghataora, Suneal, Richard M Smith, Modestos Athanasiou, and Tao Wang. 2018. “Electrically Injected Hybrid Organic/Inorganic III-Nitride White Light-Emitting Diodes with Nonradiative Förster Resonance Energy Transfer.” ACS Photonics 5 (2): 642–647. doi:10.1021/acsphotonics.7b01291.

32

Optoelectronic Organic-Inorganic Semiconductor Heterojunctions

Giordano, Anthony J, Federico Pulvirenti, Talha M Khan, Canek Fuentes-Hernandez, Karttikay Moudgil, Jared H Delcamp, Bernard Kippelen, Stephen Barlow, and Seth R Marder. 2015. “Organometallic Dimers: Application to Work-Function Reduction of Conducting Oxides.” ACS Applied Materials & Interfaces 7 (7): 4320. doi:10.1021/am5087648. Greiner, Mark T, Michael G Helander, Wing Man Tang, Zhi Bin Wang, Jacky Qiu, and Zheng Hong Lu. 2012. “Universal Energy-Level Alignment of Molecules on Metal Oxides.” Nature Materials 11 (1): 76–81. doi:10.1038/nmat3159. Guo, Song, Sang Bok Kim, Swagat K Mohapatra, Yabing Qi, Tissa Sajoto, Antoine Kahn, Seth R Marder, and Stephen Barlow. 2012a. “N-Doping of Organic Electronic Materials Using Air-Stable Organometallics.” Advanced Materials 24 (5): 699. doi:10.1002/ adma.201103238. Guo, Song, Swagat K Mohapatra, Alexander Romanov, Tatiana V Timofeeva, Kenneth I Hardcastle, Kada Yesudas, Chad Risko, Jean-Luc Brédas, Seth R Marder, and Stephen Barlow. 2012b. “N-Doping of Organic Electronic Materials Using Air-Stable Organometallics: A Mechanistic Study of Reduction by Dimeric Sandwich Compounds.” Chemistry—A European Journal 18 (46): 14760. doi:10.1002/chem.201202591. Ha, Ye Eun, Mi Young Jo, Juyun Park, Yong-Cheol Kang, Seong Il Yoo, and Joo Hyun Kim. 2013. “Inverted Type Polymer Solar Cells with Self-Assembled Mono Layer Treated ZnO.” The Journal of Physical Chemistry C 117 (6): 2646. doi:10.1021/jp311148d. Hotchkiss, Peter J, Simon C Jones, Sergio A Paniagua, Asha Sharma, Bernard Kippelen, Neal R Armstrong, and Seth R Marder. 2012. “The Modification of Indium Tin Oxide with Phosphonic Acids: Mechanism of Binding, Tuning of Surface Properties, and Potential for Use in Organic Electronic Applications.” Accounts of Chemical Research 45 (3): 337. doi:10.1021/ar200119g. Hotchkiss, Peter J, Michal Malicki, Anthony J Giordano, Neal R Armstrong, and Seth R Marder. 2011. “Characterization of Phosphonic Acid Binding to Zinc Oxide.” The Journal of Materials Chemistry 21 (9): 3107. doi:10.1039/C0JM02829K. Hwang, Jaehyung, Alan Wan, and Antoine Kahn. 2009. “Energetics of Metal-Organic Interfaces: New Experiments and Assessment of the Field.” Materials Science and Engineering R: Reports. doi:10.1016/j.mser.2008.12.001. Itskos, G, C R Belton, G Heliotis, I M Watson, M D Dawson, R Murray, and D D C Bradley. 2009. “White Light Emission via Cascade Förster Energy Transfer in (Ga, In)N Quantum Well/Polymer Blend Hybrid Structures.” Nanotechnology 20 (27): 275207. Itskos, G, X Xristodoulou, E Iliopoulos, S Ladas, S Kennou, M Neophytou, and S Choulis. 2013. “Electronic and Interface Properties of Polyfluorene Films on GaN for Hybrid Optoelectronic Applications.” Applied Physics Letters 102 (6): 063303. doi:10.1063/1.4792211. Kedem, Nir, Sylke Blumstengel, Fritz Henneberger, Hagai Cohen, Gary Hodes, and David Cahen. 2014. “Morphology-, Synthesis- and Doping-Independent Tuning of ZnO Work Function Using Phenylphosphonates.” Physical Chemistry Chemical Physics 16 (18): 8310. doi:10.1039/C3CP55083D. Kim, Yong Hyun, Jin Soo Kim, Won Mok Kim, Tae Yeon Seong, Jonghee Lee, Lars MüllerMeskamp, and Karl Leo. 2013. “Realizing the Potential of ZnO with Alternative NonMetallic Co-Dopants as Electrode Materials for Small Molecule Optoelectronic Devices.” Advanced Functional Materials 23 (29): 3645–3652. doi:10.1002/ adfm.201202799. Kobin, Björn, Jutta Schwarz, Beatrice Braun-Cula, Moritz Eyer, Anton Zykov, Stefan Kowarik, Sylke Blumstengel, and Stefan Hecht. 2017. “Spiro-Bridged Ladder-Type Oligo(ParaPhenylene)s: Fine Tuning Solid State Structure and Optical Properties.” Advanced Functional Materials 27 (45): 1–10. doi:10.1002/adfm.201704077. Koch, N, A Kahn, J Ghijsen, J J Pireaux, J Schwartz, R L Johnson, and A Elschner. 2003. “Conjugated Organic Molecules on Metal versus Polymer Electrodes: Demonstration of

Energy-Level Alignment at Organic–Inorganic Heterojunctions

33

a Key Energy Level Alignment Mechanism.” Applied Physics Letters 82 (1): 70–72. doi:10.1063/1.1532102. Koch, N, S Duhm, J P Rabe, A Vollmer, and R L Johnson. 2005a. “Optimized Hole Injection with Strong Electron Acceptors at Organic-Metal Interfaces.” Physical Review Letters 95 (23): 237601. doi:10.1103/PhysRevLett.95.237601. Koch, Norbert, Georg Heimel, Jishan Wu, Egbert Zojer, Robert L Johnson, Jean Luc Brédas, Klaus Müllen, and Jürgen P Rabe. 2005b. “Influence of Molecular Conformation on Organic/Metal Interface Energetics.” Chemical Physics Letters 413 (4–6): 390–395. doi:10.1016/j.cplett.2005.08.004. Lange, Ilja, Sina Reiter, Michael Pätzel, Anton Zykov, Alexei Nefedov, Jana Hildebrandt, Stefan Hecht, et al. 2014. “Tuning the Work Function of Polar Zinc Oxide Surfaces Using Modified Phosphonic Acid Self-Assembled Monolayers.” Advanced Functional Materials 24 (44): 7014. doi:10.1002/adfm.201401493. Mao, Hong Ying, Fabio Bussolotti, Dong Chen Qi, Rui Wang, Satoshi Kera, Nobuo Ueno, Andrew Thye Shen Wee, and Wei Chen. 2011. “Mechanism of the Fermi Level Pinning at Organic Donor-Acceptor Heterojunction Interfaces.” Organic Electronics 12 (3): 534–540. doi:10.1016/j.orgel.2011.01.003. Mcintyre, D E Aspnes 1971. “Differential Reflection Spectroscopy of Very Thin Surface Films.” Surface Science 24: 417–434. Meisel, Tino, Mino Sparenberg, Marcel Gawek, Sergey Sadofev, Björn Kobin, Lutz Grubert, Stefan Hecht, Emil List-Kratochvil, and Sylke Blumstengel. 2018. “Fingerprint of Charge Redistribution in the Optical Spectra of Hybrid Inorganic/Organic Semiconductor Interfaces.” Journal of Physical Chemistry C 122 (24): 12913–12919. doi:10.1021/acs. jpcc.8b03580. Niederhausen, J, P Amsalem, J Frisch, A Wilke, A Vollmer, R Rieger, K Müllen, J P Rabe, and N Koch. 2011. “Tuning Hole-Injection Barriers at Organic/Metal Interfaces Exploiting the Orientation of a Molecular Acceptor Interlayer.” Physical Review B - Condensed Matter and Materials Physics 84 (16): 165302. doi:10.1103/PhysRevB.84.165302. Olthof, Selina, Shafigh Mehraeen, Swagat K. Mohapatra, Stephen Barlow, Veaceslav Coropceanu, Jean Luc Brédas, Seth R. Marder, and Antoine Kahn. 2012. “Ultralow Doping in Organic Semiconductors: Evidence of Trap Filling.” Physical Review Letters 109 (17): 176601. doi:10.1103/PhysRevLett.109.176601. Oosterhout, Stefan D, L Jan Anton Koster, Svetlana S van Bavel, Joachim Loos, Ole Stenzel, Ralf Thiedmann, Volker Schmidt, et al. 2011. “Controlling the Morphology and Efficiency of Hybrid ZnO:Polythiophene Solar Cells Via Side Chain Functionalization.” Advanced Functional Materials 1 (1): 90. doi:10.1002/aenm.201000022. Ostapenko, Alexandra, Tobias Klöffel, Jens Eußner, Klaus Harms, Stefanie Dehnen, Bernd Meyer, and Gregor Witte. 2016. “Etching of Crystalline ZnO Surfaces upon Phosphonic Acid Adsorption: Guidelines for the Realization of Well-Engineered Functional SelfAssembled Monolayers.” ACS Applied Materials and Interfaces 8 (21): 13472–13483. doi:10.1021/acsami.6b02190. Piersimoni, Fortunato, Raphael Schlesinger, Johannes Benduhn, Donato Spoltore, Sina Reiter, Ilja Lange, Norbert Koch, Koen Vandewal, and Dieter Neher. 2015. “Charge Transfer Absorption and Emission at ZnO/Organic Interfaces.” Journal of Physical Chemistry Letters 6 (3): 500–504. doi:10.1021/jz502657z. Proehl, Holger, Robert Nitsche, Thomas Dienel, Karl Leo, and Torsten Fritz. 2005. “In Situ Differential Reflectance Spectroscopy of Thin Crystalline Films of PTCDA on Different Substrates.” Physical Review B - Condensed Matter and Materials Physics 71 (16): 1– 14. doi:10.1103/PhysRevB.71.165207. Rangger, Gerold M, Oliver T Hofmann, Lorenz Romaner, Georg Heimel, Benjamin Bröker, Ralf Peter Blum, Robert L Johnson, Norbert Koch, and Egbert Zojer. 2009. “F4TCNQ on Cu, Ag, and Au as Prototypical Example for a Strong Organic Acceptor on Coinage

34

Optoelectronic Organic-Inorganic Semiconductor Heterojunctions

Metals.” Physical Review B - Condensed Matter and Materials Physics 79 (16): 165306. doi:10.1103/PhysRevB.79.165306. Romaner, Lorenz, Georg Heimel, Jean Luc Brédas, Alexander Gerlach, Frank Schreiber, Robert L. Johnson, Jörg Zegenhagen, Steffen Duhm, Norbert Koch, and Egbert Zojer. 2007. “Impact of Bidirectional Charge Transfer and Molecular Distortions on the Electronic Structure of a Metal-Organic Interface.” Physical Review Letters 99 (25): 256801. doi:10.1103/PhysRevLett.99.256801. Schlesinger, R, F Bianchi, S Blumstengel, C Christodoulou, R Ovsyannikov, B Kobin, K Moudgil, et al. 2015. “Efficient Light Emission from Inorganic and Organic Semiconductor Hybrid Structures by Energy-Level Tuning.” Nature Communications 6 (1): 6754. doi:10.1038/ncomms7754. Schlesinger, Raphael, Fabio Bussolotti, Jinpeng Yang, Sergey Sadofev, Antje Vollmer, Sylke Blumstengel, Satoshi Kera, Nobuo Ueno, and Norbert Koch. 2019a. “Gap States Induce Soft Fermi Level Pinning upon Charge Transfer at ZnO/Molecular Acceptor Interfaces.” Physical Review Materials 3 (7): 074601. doi:10.1103/PhysRevMaterials.3.074601. Schlesinger, Raphael, Stefanie Winkler, Matthias Brandt, Sylke Blumstengel, Ruslan Ovsyannikov, Antje Vollmer, and Norbert Koch. 2019b. “Energy Level Alignment at Organic/Inorganic Semiconductor Heterojunctions: Fermi Level Pinning at the Molecular Interlayer with a Reduced Energy Gap.” Physical Chemistry Chemical Physics 21 (27): 15072–15079. doi:10.1039/c9cp02763g. Schlesinger, Raphael, Yong Xu, Oliver T. Hofmann, Stefanie Winkler, Johannes Frisch, Jens Niederhausen, Antje Vollmer, et al. 2013. “Controlling the Work Function of ZnO and the Energy-Level Alignment at the Interface to Organic Semiconductors with a Molecular Electron Acceptor.” Physical Review B - Condensed Matter and Materials Physics 87 (15): 155311. doi:10.1103/PhysRevB.87.155311. Schöttner, L, S Erker, R Schlesinger, Koch, A Nefedov, O T Hofmann, and C Wöll. 2020. “Doping-Induced Electron Transfer at Organic/Oxide Interfaces: Direct Evidence from Infrared Spectroscopy.” Journal of Physical Chemistry C 124 (8): 4511–4516. doi:10.1021/acs.jpcc.9b08768. Schultz, T, R Schlesinger, Niederhausen, F Henneberger, S Sadofev, S Blumstengel, A Vollmer, et al. 2016. “Tuning the Work Function of GaN with Organic Molecular Acceptors.” Physical Review B 93 (12): 125309. doi:10.1103/PhysRevB.93.125309. Koh Sharon E, Krystal D McDonald, David H Holt, and Charles S Dulcey, John A Chaney, and and Pehr E Pehrsson. 2006. “Phenylphosphonic Acid Functionalization of Indium Tin Oxide: Surface Chemistry and Work Functions.” Langmuir 22 (14): 6249. doi:10.1021/ la052379e. Smoluchowski, R 1941. “Anisotropy of the Electronic Work Function of Metals.” Physical Review 60 (9): 661–674. doi:10.1103/PhysRev.60.661. Sze, Simon M, and Daniel C Mattis. 1970. “Physics of Semiconductor Devices.” Physics Today 23 (6): 75–75. doi:10.1063/1.3022205. Timpel, Melanie, Marco V Nardi, Stefan Krause, Giovanni Ligorio, Christos Christodoulou, Luca Pasquali, Angelo Giglia, et al. 2014. “Surface Modification of ZnO(0001)-Zn with Phosphonate-Based Self-Assembled Monolayers: Binding Modes, Orientation, and Work Function.” Chemistry of Materials 26 (17): 5042–5050. doi:10.1021/cm502171m. Timpel, Melanie, Marco V Nardi, Giovanni Ligorio, Berthold Wegner, Michael Pätzel, Björn Kobin, Stefan Hecht, and Norbert Koch. 2015. “Energy-Level Engineering at ZnO/ Oligophenylene Interfaces with Phosphonate-Based Self-Assembled Monolayers.” ACS Applied Materials and Interfaces 7 (22): 11900–11907. doi:10.1021/acsami.5b01669. Vilan, A, A Shanzer, and D Cahen. 2000. “Molecular Control over Au/GaAs Diodes.” Nature 404 (6774): 166.

Energy-Level Alignment at Organic–Inorganic Heterojunctions

35

Vilan, Ayelet, Omer Yaffe, Ariel Biller, Adi Salomon, Antoine Kahn, and David Cahen. 2010. “Molecules on Si: Electronics with Chemistry.” Advanced Materials 22 (2): 140. doi:10.1002/adma.200901834. Wang, Haiyuan, Sergey V Levchenko, Thorsten Schultz, Norbert Koch, Matthias Scheffler, and Mariana Rossi. 2019. “Modulation of the Work Function by the Atomic Structure of Strong Organic Electron Acceptors on H-Si(111).” Advanced Electronic Materials 5 (5): 1800891. doi:10.1002/aelm.201800891. Winkler, Stefanie, Johannes Frisch, Raphael Schlesinger, Martin Oehzelt, Ralph Rieger, Joachim Räder, Jürgen P Rabe, Klaus Müllen, and Norbert Koch. 2013. “The Impact of Local Work Function Variations on Fermi Level Pinning of Organic Semiconductors.” Journal of Physical Chemistry C 117 (43): 22285–22289. doi:10.1021/jp401919z. Yang, Jin-Peng Peng, Lin-Tai Tai Shang, Fabio Bussolotti, Li-Wen Wen Cheng, Wen-Qing Qing Wang, Xiang-Hua Hua Zeng, Satoshi Kera, Yan-Qing Qing Li, Jian-Xin Xin Tang, and Nobuo Ueno. 2017. “Fermi-Level Pinning Appears upon Weak Electrode-Organic Contact without Gap States: A Universal Phenomenon.” Organic Electronics 48 (September): 172–178. doi:10.1016/j.orgel.2017.06.005.

3

Molecular Layer Deposition of Organic– Inorganic Hybrid Materials Xiangbo Meng

CONTENTS 3.1 Introduction...................................................................................................... 37 3.2 The Basics of MLD......................................................................................... 38 3.2.1 Surface Chemistry������������������������������������������������������������������������������ 38 3.2.2 Growth Characteristics����������������������������������������������������������������������� 41 3.3 MLD Processes for Organic–Inorganic Hybrid Metalcones........................... 41 3.3.1 Alucones��������������������������������������������������������������������������������������������� 41 3.3.1.1 Homobifunctional Organic Precursors��������������������������������41 3.3.1.2 Heterobifunctional Organic Precursors�������������������������������� 47 3.3.2 Titanicones������������������������������������������������������������������������������������������ 50 3.3.3 Zincones���������������������������������������������������������������������������������������������� 52 3.3.4 Other Metalcones�������������������������������������������������������������������������������� 54 3.4 Other Hybrid Materials.................................................................................... 56 3.4.1 Luminescent Hybrid Materials����������������������������������������������������������� 56 3.4.2 Metal-Organic Frameworks (MOFs)�������������������������������������������������� 56 3.4.3 Energy-Storage Materials������������������������������������������������������������������� 59 3.4.4 Organic Magnets��������������������������������������������������������������������������������� 60 3.4.5 Complex MLD Processes������������������������������������������������������������������� 60 3.4.6 Organic–Inorganic Hybrid Nanolaminates by MLD and ALD���������� 61 3.5 Conclusions...................................................................................................... 62 Acknowledgements................................................................................................... 62 References................................................................................................................. 62

3.1 INTRODUCTION Nanoscience and nanotechnology are playing an ever-increasing importance in our society. Scientists and engineers have been striving to manipulate atoms and molecules precisely at will to develop ideal nanosized materials with exceptional properties.1 To this end, various techniques have been devised for nanofabrication, such as mechanochemistry,2,3 wet chemistry,4 physical vapor deposition (PVD),5 chemical vapor deposition (CVD),6 atomic layer deposition (ALD),7 and molecular layer deposition (MLD).8 Among all the methods, recently MLD has been attracting more and more attention for growing pure polymeric and hybrid films.8–10 37

38

Optoelectronic Organic–Inorganic Semiconductor Heterojunctions

MLD was first coined in 1991 by Yoshimura and co-workers11 exclusively for nanoscale films of organic materials, especially pure polymers and metal-based hybrid polymers.9,12–15 This unique controllable technique was first demonstrated for synthesizing polyimides.11 Subsequently, more polymeric films were developed via MLD, ­including polyazomethines,16–19 polyureas,20–24 polyamides,25–28 poly (3,4-ethylenedi­ oxythiophene)29,30 polyimide-polyamides,31 polythioureas,32 polyethylene terephthalate,33 and some others.34,35 It was in 2008 when the first metal-based hybrid polymer was reported, which was an aluminum alkoxide (the so-called “alucone”).36 Thereafter, many more alucones13,37–45 have been developed by MLD and also ignited research interests on other metalcones, resulting in mangancone,45 zincones,46–53 zircones,54,55 titanicones,56–59 hafnicones,60 and vanadicone.61 This greatly extended our capabilities in searching for advanced materials in a controllable mode. In this regard, some excellent review papers8–10,12 have well documented MLD processes and their capabilities. Owing to its unlimited possibilities for new nanoscale polymeric films, MLD has exhibited great potentials for a large variety of applications, such as microelectronics,62 catalysis,63 energy conversion and storage,64 organic magnets,62 luminescent devices,14 surface engineering,10,65 and many others.15 Recently, there has been an increasing interest in MLD polymeric and hybrid films for addressing issues in rechargeable batteries. In this context, organic–inorganic hybrid materials are particularly intriguing, ascribed to their desirable properties unattainable with the conventional materials. This chapter focuses on introducing recent MLD research progresses on organic– inorganic hybrid materials, featuring their surface chemistry, growth characteristics, and film properties. Following this introductory section, we present some general basics commonly shared by MLD processes, including surface chemistry and growth characteristics, compared to those of ALD processes. The third part summarizes MLD processes of metalcones and the fourth part gives an account of other hybrid materials. In the last part, we conclude this chapter and give some outlook on future studies.

3.2 THE BASICS OF MLD 3.2.1 Surface Chemistry MLD and ALD are two highly similar vapor-phase techniques for nanofabrication. They share the same operational principle to realize accurate controls over materials growth. They both commonly rely on alternative self-limiting surface reactions for materials growth. The former produces organic materials while the latter results in inorganic materials. They both grow materials accurately in a layer-by-layer mode. Figure 3.1a illustrates an ALD process for growing binary inorganic materials, while Figure 3.1b displays an MLD process for growing pure polymers using two homobifunctional precursors. In terms of surface reactions, for example, the model ALD process of Al2O3 using trimethylaluminum (TMA, Al(CH3)3) and H2O can be described in Equations 3.1A and 3.1B as follows66:

OH  Al  CH 3 3  g   O  Al  CH 3 2  CH 4  g 



O  Al  CH 3 2  2H 2O  g   OAl  OH 2  2CH 4  g 



(3.1A) (3.1B)

Molecular Layer Deposition of Organic–Inorganic Hybrid Materials

39

FIGURE 3.1  Illustrations of (a) ALD and MLD processes for growing (b) pure polymeric films and (c) organic–inorganic hybrid films.

where “|” indicates substrate surfaces while “(g)” signifies gas phases. The surface chemistry of the ALD Al2O3 is based on ligand exchanges between –OH and –CH3 to arrange atoms accurately in a layer-by-layer mechanism. In addition to the ligand exchange mechanism as illustrated in Equations 3.1A and 3.1B, there are other mechanisms for ALD surface chemistry as well, such as dissociation and association.67 The four steps in Figure 3.1a constitute one ALD cycle and they can repeat to build up films for desired thicknesses. The growth rate of ALD is described by growth per cycle (GPC), typically having a GPC of ~ 1 Å/cycle.12 In the case of MLD of pure polymers (Figure 3.1b), one precursor first reacts with surface reactive groups via a corresponding linking chemistry to add a molecular layer on the substrate surface with new reactive sites.62 Following a thorough purge, another precursor reacts with the new reactive sites with the production of another molecular layer and recovers the surface back to the initial reactive groups. Another full purge is performed to finish one MLD cycle. Through repeating the afore-discussed four steps, MLD can realize polymeric film growth accurately at the molecular level. The growth rate of MLD also is described by GPC. Using adipoyl chloride (AC) and 1,6-hexanediamine (HD) as precursors,25,26 for example, an MLD process has been developed for growing nylon films linearly and the surface chemistry is described as follows:

NH 2  ClCO  CH 2 4 COCl  g   NHCO  CH 2 4 COCl  HCl  g  (3.2A) NHCO  CH 2 4 COCl  H 2 N  CH 2 6 NH 2  g 



 NHCO  CH 2 4 CO  NH  CH 2 6 NH 2  HCl  g 

(3.2B)

The AC-HD MLD process could realize a GPC of 19 Å/cycle at 62 °C.26 It is apparent that the molecular layers of –CO(CH2)4CO– and –NH(CH2)6NH– during the

40

Optoelectronic Organic–Inorganic Semiconductor Heterojunctions

MLD-nylon are much larger than the atomic layers of –Al– and –O– in the ALD-Al2O3. This underlies the higher GPC of the MLD nylon process of AC-HD. Additionally, many more pure polymeric materials via MLD recently have been summarized in literature.8 In addition to fabricating pure polymeric films as illustrated in Figure 3.1b, MLD also enables organic–inorganic hybrid materials by adopting an ALD precursor and an MLD precursor (Figure 3.1c), such as metal alkoxide materials (i.e., metalcones), in which diols can be used to couple with a metal precursor. Using TMA and ethylene glycol (EG, HOCH2CH2OH, a homobifunctional diol precursor), for instance, George’s group first reported a metal-based hybrid polymer, an aluminum alkoxide (i.e., alucone) of Al(OCH2CH2O)2 with the following surface chemistry36: OH  Al  CH 3 3  g   O  Al  CH 3 2  CH 4  g  (3.3A) O  Al  CH   2HOCH CH OH  g   OAl  OCH CH OH   2CH  g  3 2 2 2 2 2 4 2  (3.3B)  Apparently, the molecular fragment of –OCH2CH2O– attached in the MLD-alucone is far much larger than the atomic part of –O– in the ALD-Al2O3. This well explains that the resultant alucone has grown much faster than the ALD Al2O3, accounting for 4 Å/cycle at 85 °C for the MLD alkoxide36 versus 1.3 Å/cycle for the ALD Al2O3 at 80 °C.66 Through smartly selecting precursors for their functional groups and backbones, MLD enables different metalcones or hybrid materials with desired properties. Substituting EG with the aromatic hydroquinone (HQ, HOC6H4OH), for example, another alucone has been deposited and has its surface chemistry as follows38:

OH  Al  CH 3 3  g  O  Al  CH 3 2  CH 4  g 

(3.4A)

O  Al  CH 3 2  2HOC6H 4OH  g  OAl  OC6H 4OH 2  2CH 4  g  (3.4B)

This TMA-HQ MLD process exhibits a GPC of 4.1 Å/cycle at 150 °C.38 Alucones with different backbones are expected to exhibit different properties. The aromatic backbone of HQ is expected to provide structural stability and contribute largely to the electrical properties of the resultant polymer films. To date, many more metalcones have been reported, including alucones,13,36–45,53,65,68–84 zincones,47–53,85,86 titanicones,46–53,59,85,86 vanadicones,61 zircones,54,55 hafnicones,60 mangancones,45 metal quinolones,87,88 and some other hybrid materials.89–104 To investigate the underlying mechanism of the many MLD processes, a suite of in situ techniques have been employed in previous studies. Fourier transform infrared spectroscopy (FTIR)24,36,38,41–44,78 and quartz crystal microbalance (QCM)36,38,39,42–44,54,56,93 are two widely utilized in situ instruments. They are very helpful to get insightful information on surface chemistry of MLD processes. FTIR spectra can clearly identify a molecular fingerprint after each surface reaction. QCM can definitively detect any molecular deposition in mass uptake and demonstrate a linear growth for any feasible MLD processes. On the other hand, quadrupole mass spectrometry (QMS)69 is also very useful to detect any byproducts resulted from MLD surface reactions. All the data collected by the three in situ tools jointly help construct the underlying surface chemistry during a MLD process.

Molecular Layer Deposition of Organic–Inorganic Hybrid Materials

41

3.2.2 Growth Characteristics Like ALD processes, MLD processes also are subject to three parameters, that is, precursor, temperature, and substrate. MLD precursors should be able to produce sufficient vapors easily. They should also be chemically stable at deposition temperatures and highly reactive to their coupled precursors. As discussed above using EG and HQ for alucones, MLD films are highly related to the precursors adopted, such as their structures, properties, and GPCs. On the other hand, substrates have impacts on MLD growth. In some cases, substrates should be pretreated for functionalization in order to initiate the growth of some polymeric films. On the contrary, sometimes substrates should be pretreated with a protective layer in order to resist any film growth.105,106 Furthermore, deposition temperature often is critical for MLD film growth and it is worth noting that most of the MLD processes reported to date show a decreasing growth tendency with temperature. Ascribed to its unique growth mechanism, MLD produces uniform and conformal coatings over any shaped substrates.

3.3 MLD PROCESSES FOR ORGANIC–INORGANIC HYBRID METALCONES Hybrid materials are very promising in a large variety of applications such as optics, electronics, mechanics, membranes, new energies, catalysis, and surface engineering. As recently summarized in our review article,8 MLD has fabricated a variety of hybrid materials through adopting one typical ALD metal-containing precursor as the metal source and one MLD organic precursor. In addition, MLD processes can proceed with multiple precursors as well. The metal-containing ALD precursors have been collected in Figure 3.2 while the organic precursors are summarized in Figures 3.3 and 3.4 for growing hybrid materials. In this chapter, we focus on discussing MLD processes for growing metalcones. In terms of metal elements, there to date have been reported seven types of MLD metalcones, including alucones, titanicone, zincone, zircone, hafnicone, mangancone, and vanadicone.

3.3.1 Alucones Alucones are polymeric aluminum alkoxide materials with carbon-containing backbones, that is, ···Al-O-R-O-Al···, rather than the Al-O-Al backbone associated with alumoxanes.107 This type of polymer was first reported in solution-based methods by Schlenker in 1958.107 In producing alucone using MLD, TMA is the predominant metal precursor while there are many choices for an organic precursor (Figures 3.5 and 3.6). These organic precursors can be divided into two classes: homobifunctional (i.e., EG, PPDA, TEA, LC, GL, HQ, BDO, and HDO) and heterobifunctional (i.e., EA and MA, GLY, and LAC) reactants. These two types of MLD precursors showed some distinct impacts on film growth characteristics. 3.3.1.1 Homobifunctional Organic Precursors The first MAD alucone was reported through coupling TMA and EG (see Figure 3.5, AC-1).36 The resultant alucone showed a temperature-dependent decreasing

42

Optoelectronic Organic–Inorganic Semiconductor Heterojunctions

FIGURE 3.2  Metal-containing MLD precursors for organic–inorganic hybrid materials: (a) halides and (b) metal-organic compounds.

Molecular Layer Deposition of Organic–Inorganic Hybrid Materials

43

FIGURE 3.3  Homobifunctional MLD precursors for organic–inorganic hybrid films, featuring the reactive groups: (a) diol, (b) amine, (c) thiol, chloride, isocyanate, and uracil.

tendency in GPC from 4 Å/cycle at 85 °C to 0.4 Å/cycle at 175 °C. The surface reactions are shown in Equations of 3.3A and 3.3B. Using FTIR and QCM, Dameron et  al.36 also disclosed an alternative growth mechanism for this alucone MLD. At higher temperatures (e.g., 135 °C), there are no noticeable O–H stretching vibrations observed, suggesting the EG molecules might have reacted twice with –AlCH3 species. Thus, it is believed that, during TMA-EG deposition, TMA molecules might have diffused into the alucone polymer film while reacted with –OH species. The EG

44

Optoelectronic Organic–Inorganic Semiconductor Heterojunctions

FIGURE 3.4  Other MLD precursors for organic–inorganic hybrid films, featuring the reactive groups: (a) three groups, (b) ethylene, (c-f) heterofunctional groups.

molecules might then react with –AlCH3 species both on the alucone polymer surface and on TMA molecules in the alucone film. Dameron et al.36 also revealed that the resultant MLD alucone films are not stable in air and their thickness decreases with time over the first 150 hours after their fabrication. At the meantime, the films’ composition changes, due to either dehydration or dehydrogenation reactions.

Molecular Layer Deposition of Organic–Inorganic Hybrid Materials

45

FIGURE 3.5  MLD processes for alucones by coupling TMA with homobifunctional precursors.

46

Optoelectronic Organic–Inorganic Semiconductor Heterojunctions

FIGURE 3.6  MLD processes for alucones by coupling TMA with heterofunctional precursors.

Similarly, BDO (see Figure 3.5, AC-2) and HDO (see Figure 3.5, AC-3) also were used to couple with TMA for growing alucones.37 Park et al.37 revealed that, due to the longer carbon chains of BDO and HDO than that of EG, they are prone to cause double reactions, that is, the two ends of EG both reacted with TMA species (–Al(CH3)2). Consequently, the resultant MLD films are more possible with holes formed by the double reactions. Different from the flexible chain structure of EG, HQ is a homobifunctional diol but has a rigid structure. Choudhury and Sarkar38 studied the MLD growth of TMA-HQ (see Figure 3.5, AC-4) using in situ FTIR and QCM and disclosed a linear growth in the range of 150–225 °C with a GPC of 4.1 Å/cycle at 150 °C and 3.5 Å/ cycle at 225 °C. They also revealed that the TMA-HQ films are prone to degrade in air, but an ALD-Al2O3 capping layer can help improve their stability. With the same backbone as FQ, FHQ (Figure 3.5, AC-5) was also reported but the related MLD was not well studied.45 In addition, PPDA has the same backbone as HQ but features two amines instead of hydroxyls. Zhou et al.13 studied the TMA-PPDA growth (Figure 3.5, AC-9) and revealed that the MLD process of TMA-PPDA exhibited a GPC of 1.4 Å/cycle at 400 °C. In addition, the resultant TMA-PPDA films are airsensitive and show a severe increase by 30% in thickness when exposed to air for 2 weeks.13 Given the poor stability of the TMA-PPDA films in air, Zhou alloyed this alucone with ALD-Al2O3 and the resultant nanolaminates exhibited improved stability in air.13 In addition, the 1:4 alucone/Al2O3 nanolaminates showed tunable electrical properties.13 TEA as a homobifunctional precursor was first preliminarily studied by Bahlawane et al.,39 but was investigated with more efforts by Lemaire et al. (see Figure 3.5, AC-8).79 It was revealed that the resultant amine-containing alucone showed a decreasing growth tendency from 6.7 Å/cycle at 150 °C to 0.8 Å/cycle at 195 °C.

Molecular Layer Deposition of Organic–Inorganic Hybrid Materials

47

In  addition to the organic precursors discussed above, there are also some other homofunctional precursors exposed in literature, that is, LC (Figure 3.5, AC-6)68 and GL (Figure 3.5, AC-7).81 3.3.1.2 Heterobifunctional Organic Precursors Homobifunctional organic reactants (such as EG, BDO, and HDO) typically undergo a symmetric “double-end” surface reaction. This parasitic reaction process decreases the density of reactive sites available for the subsequent half-reaction and results in slow growth rates and poor material stability.42 To this end, heterobifunctional reactants and ring-open reaction were suggested.12 In this regard, Yoon et al.41 first reported an ABC MLD process using TMA, EA, and MA as precursors, in which a heterobifunctional precursor EA and a ring-open reaction of MA were employed, as illustrated in Figure 3.7. The related surface chemistry could also be described as follows41: OH  Al  CH 3 3  g   O  Al  CH 3 2  CH 4  g  (3.5A) O  Al  CH 3   2NH 2CH 2CH 2OH  g   OAl  OCH 2CH 2 NH 2   2CH 4  g  2 2  (3.5B)  OAl  OCH 2CH 2 NH 2 2  C4H 2O3  g   OAl (3.5C)  OCH2CH2NH  COCHCHCOOH 2  g  One end of the heterobifunctional reactant might react preferentially to avoid the double reaction. Likewise, ring-open reactants might react and yield a new functional group upon ring-opening that did not react with the initial surface species. In the temperature range of 90–170 °C, Yoon et al.41 monitored this three-step alucone film using in situ FTIR and found that the three sequential surface reactions displayed self-limiting growth, showing a decreasing growth trend with temperature from 24 Å/ cycle at 90 °C to 4.0 Å/cycle at 170 °C. Using QCM in a subsequent study, the team revealed that there was a significant diffusion of TMA into the afore-deposited ABC films, contributing to an extraordinary mass gain.82 It seemed that the pre-deposited ABC films have acted as a TMA reservoir during the TMA exposures. It was found that there were three regions for the three-step ABC film growth. The first region (~10 cycles on Al2O3 surface) was the formation region, in which the ABC film first reached a threshold thickness. In the second region, then, the ABC film growth showed an increase in mass gain per cycle (about another 20 cycles on Al2O3 surface). Once the ABC film was thicker than the TMA diffusion distance, the GPC enabled a constant steady state, that is, the third region (since the 30th cycle on Al2O3 surface). The diffused TMA molecules were believed to react with the following exposures of EA molecules, and the reaction was regarded as a CVD process. The CVD reaction most likely occurs close to the surface of the ABC film, for EA did not diffuse into the ABC film. In contrast, the third reaction of EA–MA was regarded only as a surface reaction, for MA was a much larger molecule and unable to diffuse into the ABC film. Seghete et al.82 also disclosed that extended purging times were helpful to reduce the diffused TMA in the ABC film. Increasing growth temperatures higher than 130 °C is

48

Optoelectronic Organic–Inorganic Semiconductor Heterojunctions

FIGURE 3.7  Schematic illustration of the three-step reaction sequence of the ABC alucone growth using (A) trimethylaluminum (TMA), (B) ethanolamine (EA), and (C) maleic anhydride (MA).41 Reprinted with permission from Ref. 41. Copyright (2009) American Chemical Society.

another effective route to eliminate TMA diffusion into the ABC film. Furthermore, Seghete et al.82 examined the ABC film’s stability using XRR at 90 °C, disclosing that the ABC film has the largest change in the first 50–70 hours and then no change in thickness after 300 hours. They found that, after aging in air, the ABC film thickness decreased by 5.8%, in comparison to a decrease of 20% with the TMA-EG film. In addition to the afore-discussed ABC MLD process, binary MLD processes have also been reported. Coupling GLY with TMA (see Figure 3.6, AC-10), for instance, two groups reported the same MLD process for an alucone independently almost at the same time.42,43 In one of the two studies, Gong et al.42 believed that,

Molecular Layer Deposition of Organic–Inorganic Hybrid Materials

49

during the exposure of GLY, a ring-opening/transalkylation reaction occurred to produce (Al-O-CH2-CH(CH3)-CH2-OH) or (Al-O-CH2-CH(CH2-CH3)-OH) groups. The GLY could bond through Lewis acid/base interaction and by methyl elimination to form Al-O-C- bonds. The remaining hydroxyls then reacted with TMA during the next step. Despite its heterobifunctional structure, however, GLY was not successful in preventing double reactions from occurrence. Gong et al. 42 revealed that the TMA-GLY MLD decreased from 24 Å/cycle at 90 °C to 6 Å/cycle at 150 °C. They also disclosed that the resultant films of TMA-GLY were stable in air and had little change at an annealing temperature of 100 °C for 2 hours. However, the as-grown films lost hydroxyl groups at 200 °C and most of C-H species at 300 °C, associated with a film shrinkage during the annealing process. In another independent study, in comparison, Lee et al.43 reported much lower growth rates of the TMA-GLY MLD, for example, a GPC of 1.3 Å/cycle at 125 °C. They noticed that the growth of TMAGLY was very sensitive to the purging durations and shorter purges could dramatically increase GPCs. This may underlie the different GPCs reported by Lee et al.43 and Gong et al.42. Interestingly, Lee et al.43 proposed a different mechanism responsible for the TMA-GLY MLD. They believed that the hydroxyl group of GLY reacted with AlCH3 surface species and produced alkoxy aluminum surface species (Figure 3.8). In addition, oxygen atom in the epoxy ring of GLY could coordinate to the Al atom via Lewis acid/base interactions (Figure 3.8b). This Lewis acid/base interaction weakened the C3-oxygen bond on the epoxy ring. In the subsequent exposure of TMA (Figure 3.8c), TMA could coordinate to the oxygen in the Al-O-C bond through another Lewis acid/base interaction. Methyl transferred from TMA to

FIGURE 3.8  Proposed surface chemistry for the TMA-GLY reaction assuming a 1:1 TMA/ GLY stoichiometry.43 Reprinted with permission from Ref. 43. Copyright (2011) American Chemical Society.

50

Optoelectronic Organic–Inorganic Semiconductor Heterojunctions

position C3 on the epoxy ring (Figure 3.8d). This methyl transfer then formed a 1,2-butanediolate as displayed in Figure 3.8a′. Ring-opening after methyl transfer concurrently could form a new Al-O bond between the oxygen in the original epoxy ring and the neighboring Al atom. The proposed reaction sequence could then be repeated by the second GLY exposure, as illustrated in Figure 3.8b′. Lee et al.43 also believed that Lewis acid/base interactions were important in the growth mechanism. There are two kinds of Lewis acid/base interactions: (1) one between the Al of the AlCH3 surface species and the oxygen in the epoxy ring and (2) another between the Al of the TMA reactant and the oxygen in Al-O-C bond. Lee et al.43 asserted that both the Lewis acid/base interactions were essential for the methyl transfer reaction from TMA to the epoxy ring. In particular, Lee et al.43 conducted an additional experiment using DEZ and GLY to verify the importance of Lewis acid/base interactions. They revealed that there was no film growth with the DEZ-GLY, for the Zn atom in DEZ had much less Lewis acidity that is essential for efficient epoxy ring-opening. Lee et  al.43 further confirmed that the TMA-GLY films annealed at 300 or 500 °C for 24 hours had no remaining carbon, and had turned into porous Al2O3 films of 27% porosity, in terms of the density of ALD Al2O3 films. Another heterofunctional precursor for MLD alucones is LAC (see Figure 3.6, AC-11) reported by Gong and Parsons.44 They revealed that the growth of TMA-LAC exhibited a decreasing tendency, accounting for 0.75 Å/cycle at 60 °C and 0.08 Å/cycle at 120 °C. Gong and Parsons44 explained the growth mechanism that, based on FTIR measurements, an Al-CH3 Lewis acid site could catalyze the (C=O)-O-C ring-opening to form Al-O and C-CH3 through methyl transfer from the aluminum and then the Al-O-C group was accessible for reaction during the next TMA exposure. Encouragingly, the resultant TMA-LAC films showed excellent stability in air over 30 days. In a latest work, Baek et al. used MP to couple with TMA for growing an alucone (see Figure 3.6, AC-12) in the range of 100–200 °C.108 MP is a heterofunctional organic precursor, featuring one -OH ligand and one -SH ligand on its two ends. The resultant alucone has a GPC of 2.0–2.5 Å/cycle in the temperature range. Very importantly, Baek et al. found that, different from other metalcones, this alucone are very stable in the atmospheric and humid air conditions. In addition, after annealing under vacuum at 300–750 °C, the annealed alucone films showed thermal polymerization and their carbon ring structures transformed into graphitic carbon flakes with improved electrical conductivity.

3.3.2 Titanicones Titanicones are another type of popular metalcones investigated to date. There are two Ti sources used for growing titanicones, that is, TiCl4 and TDMATi (see Figure 3.9). The organic precursors are EG, GL, TEA, and FA used in binary MLD processes (see Figure 3.9). In addition, there is a four-step ABCD MLD process for a titanicone using the sequence of TiCl4-EA-MC-EA.59 Coupling TiCl4 with EG (Figure 3.9, TC-1) and GL (Figure 3.9, TC-2), respectively, Abdulagatov et al. made a comparative study on the two MLD processes.56 They revealed that the TiCl4-EG MLD has a constant GPC of 4.5 Å/cycle in the range of 90–115 °C while the GPC decreases to 1.5 Å/cycle at 135 °C. In comparison, the

Molecular Layer Deposition of Organic–Inorganic Hybrid Materials

51

FIGURE 3.9  MLD processes for titanicones by coupling (a) TiCl4 and (b) TDMATi with organic precursors.8 Reprinted with permission from Ref. 8. Copyright (2017) The Royal Society of Chemistry.

TiCl4-GL MLD needs higher temperatures of 130–210 °C and its GPC decreases from 2.8 Å/cycle at 130 °C to 2.1 Å/cycle at 210 °C. Working on 500-nm thick films with nanoindentation, Abdulagatov et al.56 disclosed that the TiCl4-EG film has an elastic modulus of ~8 GPa and a hardness of ~0.25 GPa while the TiCl4-GL film has an elastic modulus of ~30 GPa and a hardness of ~2.62 GPa. It was further found that, compared to the TiCl4-EG film, the TiCl4-GL film has a much better thermal stability up to 250 °C, probably due to their higher network connectivity.

52

Optoelectronic Organic–Inorganic Semiconductor Heterojunctions

In another MLD process, Lemaire et al.79 used TEA to couple with TiCl4 (see Figure 3.9, TC-3) and compared the resultant titanicone with the alucone from the TMA-TEA MLD. They disclosed that, compared to the GPC of alucone that is 6.7 Å/ cycle at 150 °C and 0.8 Å/cycle at 195 °C, the resultant titanicone enables a GPC of 5.2 Å/cycle at 150 °C and 2 Å/cycle at 195 °C. Lemaire et al.79 also revealed that the TMA-TEA is more stable in ambient conditions, accounting for an increase by 5.1% after 24 hours and no more change over 500 hours versus a decrease of the TiCl4-TEA films by 6.2% after 24 hours and 14.1% after 500 hours in air. In addition, Lemaire et al.79 found that humidity has a significant impact on the TiCl4-TEA films. The film thickness increases significantly from 207 to 346.9 nm with relative humidity from 10% to 65%. Particularly, the shrinking and welling behavior is reversible and repeatable unless the relative humidity exceeds ~65%, which leads to film degradation. In contrast, the TMA-TEA film thickness has much smaller change with humidity, having a change of ~3% with relative humidity from 10% to 65%. In another work,58 Cao et al. reported TiCl4 and FA as precursor for a hybrid Ti-polymer (see Figure 3.9, TC-4), showing a decreasing GPC of 1.1 Å/cycle at 180°C and 0.49 Å/cycle at 300 °C. Alternatively, TDMATi has been used as a Ti precursor to couple with EG and GL, respectively. In a work, Van de Kerckhove et al.57 studied the resultant MLD processes for growing titanicones. The researchers found that the TDMATi-EG MLD (Figure 3.9, TC-5) terminates after a few of cycles, due to severe double reactions of EG. In comparison, the TDMATi-GL MLD process (Figure 3.9, TC-6) can work efficiently and sustain self-limiting growth in the temperature range of 80–160 °C with a linearly decreasing growth tendency from 0.95 to 0.24 Å/cycle. Van de Kerckhove et al.57 has attributed the decreasing GPCs to the aggravated desorption of TDMATi with temperature.

3.3.3 Zincones Zincones are another type of metalcones resulting from zinc reactants and organic precursors. DEZ is dominantly used as the Zn source while there have six precursors reported to date, including EG, HQ, THB, HDD, GL, and AP (Figure 3.10a). The first zincone MLD was reported by Yoon et al., using DEZ and EG as precursors (Figure 3.10a, ZC-1).46 This MLD process showed a decreasing GPC from 4 Å/ cycle at 90 °C to 0.25 Å/cycle at 170 °C. In surface chemistry, this process is similar to that of the TMA-EG and the DEZ-EG MLD processes. There is also some diffusion of DEZ into the zincone film. The diffused DEZ molecules can react with the subsequent EG exposure to form new zincone polymer chains. Furthermore, FTIR data revealed little O-H stretching vibration at 170 °C but evident O-H stretching vibration at 90 °C.46 This suggested that most of the EG molecules reacted twice to –ZnCH2CH3 at 170 °C. Particularly, XPS analyses on zincone samples showed a decreasing Zn content with temperature, changing from 10.7% at 90 °C to 10.6, 9.9, 8.9% at 110, 130, and 150 °C, respectively. Furthermore, Yoon et al.46 verified that the DEZ-EG films adsorb water after exposure to air and then become very stable for multiple weeks. The DEZ-EG MLD process has also been investigated by Peng et al.47 They confirmed a decreasing GPC with temperature in the range of 100–170 °C but reported larger GPCs. Additionally, they also verified that the films are stable in dry air for

Molecular Layer Deposition of Organic–Inorganic Hybrid Materials

53

FIGURE 3.10  MLD processes for (a) zincones and (b) zinc hybrids.8 Reprinted with permission from Ref. 8. Copyright (2017) The Royal Society of Chemistry.

3 days. Peng et al.47 also observed that the DEZ-EG films are prone to adsorb water under ambient conditions for 1 hour but have no more changes up to 12 hours. This indicates that the reaction between the DEZ-EG films and moisture from ambient environment proceeded quickly, resulting in hydrolyzed DEZ-EG films. A further annealing on the hydrolyzed DEZ-EG films at 100 °C for 2 hours produced a structure similar to the ALD ZnO. Another MLD process of zincones is DEZ-HQ. Due to the conjugated structure of HQ, the resultant DEZ-HQ was expected to produce electrically conductive

54

Optoelectronic Organic–Inorganic Semiconductor Heterojunctions

polymers.48 In addition, the rigid structure of HQ was expected to reduce the possibility of double reactions. Yoon et al. studied the DEZ-HQ MLD (see Figure 3.10a, ZC-2) in the temperature range of 130–170 °C and revealed a linear growth with a GPC of 1.6 Å/cycle at 150 °C.48 Subsequently, Yoon et al. explored conductive hybrid materials by combining ALD-ZnO with the MLD zincone of DEZ-HQ for various nanolaminates (alloys).49,50 They disclosed that, in comparison to the conductivity of 14 S/cm of ALD ZnO, the ZnO:zincone alloys of 1:1 and 2:2 cycle ratios yielded conductivity of 116 and 170 S/cm, respectively. Thus, these studies paved a venue to tune electrical conductivity of hybrid materials through combining ALD with MLD rationally. Other researchers also confirmed that the alloys of ALD-ZnO/MLDzincone are tunable in electrical properties.85,86 Other zincones include DEZ-THB (Figure 3.10a, ZC-3),51 DEZ-HDD (Figure  3.10a, ZC-4),52 DEZ-GL (Figure 3.10a, ZC-5),53 DEZ-AP (Figure 3.10a, ZC-6).96 The DEZ-THB could realize a GPC of 2.6 Å/cycle in the range of 100–160 °C and the DEZ-HDD process exhibited a GPC of 5.2 Å/cycle in the range of 100–150 °C. Particularly, the DEZ-HDD films showed excellent stability in air up to 400 °C. There also has a zinc-based hybrid material reported using DEZ-ODA (Figure 3.10b).96

3.3.4 Other Metalcones Additionally, some other metalcones have also been reported, including zircones, hafnicones, mangancones, and vanadicones (Figure 3.11). One MLD process of zircons was fulfilled using ZTB and EG as precursors (Figure 3.11a),54,55 exhibiting a decreasing GPC from 1.6 Å/cycle at 105 °C to 0.3 Å/cycle at 195 °C. Using TDMAHf as the hafnium (Hf) precursor to couple with EG, Lee et al.60 succeeded a hafnicone MLD (see Figure 3.11b), having a decreasing GPC from 1.2 Å/cycle at 105 °C to 0.4 Å/cycle at 205 °C. In addition, a MLD process of mangancones has been reported using Mn(CpEt)2 and EG as precursors (see Figure 3.11c).45 Furthermore, two processes have been reported in a study61 for vanadicones, using either EG or GL to couple with TEMAV (Figure 3.11d). In the study, it was found that the TEMAV-EG couple was not successful for growing vanadicone, possibly due to double reactions of EG. In contrast, the TEMAV-GL couple (see Figure 3.11d) enabled a reliable process for vanadicones, showing a linearly decreasing growth tendency from ­ ~1.2 Å/cycle at 80 °C to ~0.5 Å/cycle at 180 °C. More recently, Cu(dmap)2 was used to grow Cu-based metalcones (Cu-cones) with either HQ or BDC (or TPA) (see Figure 3.11e) via MLD.109 The Cu-HQ metalcone (Figure 3.11e, CC-1) was conducted in the range of 100–120 °C with a GPC of 1.0–1.5 Å/cycle while the Cu-BDC metalcone (Figure 3.11e, CC-2) was performed at 160 °C with a GPC of 2.6 Å/cycle. Furthermore, Cu(dmap)2 also has been used to couple with three other organic precursors for growing other hybrid materials via MLD in the same study. The three other organic precursors are PPDA, ODA, and BDT (see Figure 3.3). The Cu-ODA was deposited in the range of 140–220 °C with a constant GPC of 2.4 Å/cycle and the Cu-PPDA hybrid was investigated in the range of 70–200 °C with a constant GPC of 1.7 Å/cycle. The Cu-BDT was studied in the range

Molecular Layer Deposition of Organic–Inorganic Hybrid Materials

55

FIGURE 3.11  MLD processes for (a) zircones, (b) hafnicones, (c) mangancones, (d) vanadicones, and (e) Cu-cones.

56

Optoelectronic Organic–Inorganic Semiconductor Heterojunctions

of 80–140 °C, showing a decreasing GPC from 1.9 Å/cycle at 80 °C to 0.8 Å/cycle at 140 °C. All these Cu-based hybrid materials are amorphous or nanocrystalline.

3.4 OTHER HYBRID MATERIALS Besides metalcones discussed above, some other types of hybrid materials have also been reported, as summarized in Figures 3.12 and 3.13 using one metal precursor and one organic precursor. In addition, there have been some MLD processes using multiple precursors for growing hybrid materials, which provide alternative solutions for tunable fabrication of novel polymeric films, as summarized in Figure 3.14.

3.4.1 Luminescent Hybrid Materials Metal quinolones (Mqx) are very promising hybrid luminescent materials. Nilsen et al.87,110 first developed MLD processes of Mqx. Using 8-HQ to couple with TMA, DEZ, or TiCl4, Nilsen et al.87 synthesized Alq3, Znq2, or Tiq4, respectively (see Figure 3.12a). The three metal quinolones exhibit decreasing GPCs with temperature, 4–7 Å/cycle at 80 °C but no growth at 200 °C. The Alq3 was examined for stability at 85 °C and it was found that the Alq3 is stable to water at 85 °C.87 In a later study, Räupke et al. demonstrated the luminescent performance of MLD Alq3 films.88 Subsequently, another team studied luminescent hybrid materials using three metal precursors (i.e., Na(thd), Ba(thd)2, and La(thd)3) and two organic precursors (uracil and adenine).111,112 For the metal precursors, thd is 2,2,6,6-tetramethyl-3,5-heptadionate. Uracil is one of the four nucleobases (NBs) in the nucleic acid of ribonucleic acid (RNA) while adenine is one of the four NBs in the nucleic acid of deoxyribonucleic acid (DNA). Among them, the Ba(thd)2-uracil pair has been well studied and exhibited a nearly constant GPC of ~2.8 Å/cycle in the range of 260–320 °C. They found that the metal precursors have evident influence on the crystallinity of the resultant luminescent hybrid metal-NB materials. Specifically, the La-NBs are amorphous, the Ba-NBs are at least partially crystalline, and the Na-NBs are well crystalline. These metal-NB thin films were tested for their luminescent properties. Particularly, both the Na-NB and Ba-NB films show intense photoluminescence in blue and green wavelengths. This work has significant implications for MLDdeposited hybrid materials in luminescent applications. Another luminescent work was recently conducted on erbium-based hybrid materials. Using Er(DPDMG)3 (Figure 3.2) with PDA (Figure 3.3) as precursors, Mai et al. developed an MLD process for growing Er-di-3,5-pyridinecarboxylate [Er2(3,5-PDC)3] films.113 In the range of 250–265 °C, this MLD process exhibits a constant GPC of ~6.4 Å/cycle. The resultant Er2(3,5-PDC)3 films were confirmed for their promising UV absorption properties and a photoluminescence at 1535 nm for a 325-nm excitation.

3.4.2 Metal-Organic Frameworks (MOFs) Metal-organic frameworks (MOFs), also known as porous coordination polymers (PCPs), are crystalline coordination polymers built from inorganic nodes (bridging ligands) and organic linkers.114,115 One of the most remarkable features of MOFs is

Molecular Layer Deposition of Organic–Inorganic Hybrid Materials

57

FIGURE 3.12  MLD processes for (a) metal quinolones and (b) metal-organic frameworks.8 Reprinted with permission from Ref. 8. Copyright (2017) The Royal Society of Chemistry.

their extremely high porosity, consequently featuring their high surface areas and high pore volumes in uniformly tailorable pores. In addition, MOF’s mechanical and chemical properties are also tunable by carefully selecting the linker molecule or introducing guest molecules into the pores. These unique characteristics make MOFs very promising in many applications, such as gas storage and separation,116–119

58

Optoelectronic Organic–Inorganic Semiconductor Heterojunctions

FIGURE 3.13  MLD processes for (a) lithium terephthalate and (b) metal tetracyanoethylene.8 Reprinted with permission from Ref. 8. Copyright (2017) The Royal Society of Chemistry.

luminescence,120 drug delivery,121,122 energy conversion and storage,123,124 and ­catalysis.114,125,126 Salmi et al.89 initiated the first study of MLD for growing MOFs (see Figure 3.12b, MOF-5). In the work, the researchers adopted ZnAc2 and BDC as precursors for MOF-5, an isoreticular MOF (also known as IRMOF-1) having a cubic framework structure and consisting of ZnO4 clusters connected with rigid benzene dicarboxylate linkers. The resultant films exhibited a decreasing tendency in GPC from 6.5 Å/cycle at 225 °C to 4.5 Å/cycle at 325 °C. There is no film growth at 350 °C. Unfortunately, in the whole deposition temperature range, no crystalline phases have been grown. Surprisingly, however, the as-deposited amorphous films could crystalize into an unidentified structure at room temperature under a relative humidity of 60%. This moisture-induced structure was confirmed with no porosity but could be recrystallized into the MOF-5 phase in an autoclave with N,N-dimethylformamide (DMF) at 150 °C. In a subsequent study, Salmi et al. 90 used NDC to couple with ZnAc2 for growing IRMOF-8 (Figure 3.12b, IRMOF-8). This MLD of ZnAc2-NDC also exhibited a linear decreasing GPC from 4.9 Å/cycle at 260 °C to 4.0 Å/cycle at 300 °C. Again, the as-deposited films were amorphous but could be crystallized into an unknown structure at room temperature under a relative humidity of 70%. The unknown crystallized structure was further recrystallized in an autoclave with DMF at 150 °C and turned into IRMOF-8. The crystallization in humid air was confirmed being critical, for the as-deposited films could not be directly turned into IRMOF-8 only via the second crystallization. Furthermore, Salmi et al.90 verified the films via the recrystallization are porous and they successfully loaded Pd into the porous films uniformly.

Molecular Layer Deposition of Organic–Inorganic Hybrid Materials

59

FIGURE 3.14  Four-step ABCD MLD processes for organic–inorganic hybrid materials of (a) self-assembled monolayers, (b) titanicones, and (c) ZnO-polymers.8 Reprinted with permission from Ref. 8. Copyright (2017) The Royal Society of Chemistry.

A more recent work of MLD MOFs was reported by Ahvenniemi and Karppinen91 for copper(II)terephthalate (i.e., Cu-TPA, also MOF-2) using Cu(thd)2 and BDC as precursors (see Figure 3.12b, MOF-2). Very encouragingly, this MLD process of Cn(thd)2-BDC can directly grow porous crystalline Cu-TPA films in the range of 180–195 °C. However, a higher temperature of 260 °C led to amorphous films. In the whole temperature range of 180–260 °C, the GPC decreased with temperature from 3 to 0.2 Å/cycle.

3.4.3 Energy-Storage Materials Using Li(thd) and TPA as precursors, Nisula and Karppinen92 developed a lithiumcontaining crystalline hybrid polymer, LiTP, lithium terephthalate (LTP) (see Figure 3.13a). They revealed that the MLD process of LTP sustained a self-limiting growth during the temperature of 200–280 °C with a decreasing tendency from ~4.0 to 1.0 Å/cycle. Electrochemical testing confirmed that these LTP hybrid films are

60

Optoelectronic Organic–Inorganic Semiconductor Heterojunctions

promising anodes in lithium-ion microbatteries. Subsequently, this team further synthesized hybrid materials of metal-TP, using Ca(thd)2, Na(thd), K(thd), Mg(thd)2, Sr(thd)2, Ba(thd)2, and La(thd)3 to couple with TPA as precursors, respectively.127,128

3.4.4 Organic Magnets Organic molecule-based magnets (MBMs) are a relatively new class of magnetic materials and first emerged in 1985.129,130 Compared to conventional metallurgic and ceramic magnets, the main benefits of MBMs are usually associated with their lightweight, mechanical flexibility, tunable color or transparency, low-temperature processing, solubility, and compatibility with polymers and other classes of molecular materials.129 The M[TCNE] (M = 3d metal; TCNE = tetracyanoethylene) complexes represent one of the most interesting classes of MBMs, possessing numerous compositions and structures with varying dimensionalities of magnetic coupling from onedimensional (1D) inorganic polymer chains and two-dimensional (2D) layers to three-dimensional (3D) networks and amorphous solids.129,130 Kao et al. reported an MLD attempt to synthesize an MBM, V[TCNE]x using tetracyanoethylene (TCNE) and vanadium hexacarbonyl (V(CO)6) at room temperature (Figure 3.13b, MBM-1).93 V[TCNE]x is a promising MBM for the future organic/molecular electronic, magnetic, and spintronic applications. They described the MLD process of V(CO)6TCNE as follows:



  TCNE  xV  CO 6  g    TCNE  V  CO y



  TCNE  V  CO y

  x 6  y  CO  g  x

(3.6A)

  zTCNE  g    TCNE  V  TCNE  xyCO  g  (3.6B) x

x

z

Kao et al. revealed a linear growth with a GPC of 9.8 Å/cycle. In a subsequent effort, Kao et al. further synthesized two more MBMs using Co2(CO)8 as the Co source, that  is, Co[TCNE]x (see Figure 3.13b, MBM-2) and CoxVy[TCNE]z at room temperature.94

3.4.5 Complex MLD Processes Not limited to two precursors (one metal precursor and one organic precursor, as shown in Figure 3.1c), there are some MLD processes combining multiple precursors for hybrid materials. In this regard, Sung’s group first developed four-step MLD processes for growing self-assembled monolayers (SAMs), in which organic fragments are connected by metal oxides, including TOx-SAMs (Figure 3.14a, SAM-1),98 ZrOx-SAMs (Figure 3.14a, SAM-2),104 AlOx-SAMs (Figure 3.14a, SAM-3101 and SAM-4102,103). All the MOx-SAMs reported exhibited good mechanical flexibility and stability, excellent insulating properties, and relatively high dielectric capacitances with a high dielectric strength, showing promising applications as nanohybrid dielectrics.99,100,103 In addition to SAMs, Knez’s group reported an MLD process consisting of a fourstep procedure of TiCl4-EA-MC-EA to produce a titanicone (Figure 3.14b).59 Using

Molecular Layer Deposition of Organic–Inorganic Hybrid Materials

61

this coupling combination, Chen et al.59 demonstrated the high flexibility of MLD processes for a wide range of hybrid materials. At 100 °C, this four-step MLD process receives a growth rate of 6 Å/cycle. The resultant Ti-hybrid polymer films were later used for producing nitrogen-doped porous TiO2 nanotubes for improved photocatalytic activity. Additionally, Qin’s group reported another four-step MLD procedure of DEZ-PD-EDA-PD for a Zn-based hybrid polymer (Figure 3.14c)95 with a GPC of 5 Å/cycle at 100 °C.

3.4.6 Organic–Inorganic Hybrid Nanolaminates by MLD and ALD Besides the strategies discussed above, there has been another strategy combining ALD and MLD for organic–inorganic hybrid nanolaminates. Compared to MLD, ALD is a relatively well-developed technique with a large variety of inorganic materials reported to date. Thus, there is an unlimited research space for developing function-oriented nanolaminates through combining ALD and MLD. For instance, combining the MLD process of TiOx-SAMs with the ALD process of TTIP-H2O, Lee et al. demonstrated that various TiOx-SAMs/TiO2 nanolaminate films (see Figure 3.15) can be accurately realized in a highly tunable fabrication mode.98 The thickness of SAMs and TiO2 nanolayers in each sample could be controlled by adjusting the number of MLD and ALD cycles. The GPCs of the MLD SAMs and the ALD TiO2 GPC are ~11 and 0.6 Å/cycle, respectively. Similarly, the nanolaminates of AlOx-SAMs/ Al2O3, AlOx-SAMs/ZnO, AlOx-SAMs/Al2O3/AlOx-SAMs/ZnO were also reported.102

FIGURE 3.15  TEM images of self-assembled organic multilayer/TiO2 nanolaminate films fabricated using MLD and ALD (white is the SAMs layer and black is the TiO2 layer).98 Reprinted with permission from Ref. 98. Copyright (2007) American Chemical Society.

62

Optoelectronic Organic–Inorganic Semiconductor Heterojunctions

In addition, combining ALD-ZnO (using DEZ and water) and MLD processes of zinc-based polymeric films (DEZ-HQ, DEZ-AP, and DEZ-ODA), a series of nanolaminates of ZnO/zinc-polymeric film have been developed.96 Along with the increasing needs in new materials, it is expected that more and more hybrid nanolaminates will be reported with exceptional properties.

3.5 CONCLUSIONS This work focuses on summarizing the recent progresses of MLD in developing ­inorganic–organic hybrid materials. MLD exhibits highly flexible capabilities in synthesizing novel hybrid materials with exceptional properties. Firstly, MLD can develop hybrid materials through combining one ALD metal precursor with one MLD organic precursor. To date, most of the hybrid materials have been produced by this strategy. Secondly, MLD can synthesize hybrid materials through combining multiple precursors. In this regard, just a few cases have been reported. Thirdly, MLD can also combine with ALD for an unlimited amount of nanolaminates with controllable thicknesses for inorganic and organic layers. All these have determined that MLD is a very attractive technique for novel materials with desirable properties. These capabilities of MLD will enable providing new solutions to many urgent challenges in many applications. Although all these exciting progresses have been made with MLD so far, it should be noted that many more efforts are urgently needed for expanding MLD. On one hand, current MLD processes and resultant hybrid films are still unsatisfactory, mainly due to their precursors. The ideal precursors should meet high volatility, high stability, and high reactivity. In addition, the precursors should be able to avoid double-end reactions. On the other hand, the stability of organic–inorganic hybrid materials by MLD to date remains an issue. Most of MLD hybrid materials are sensitive to air, humidity, and/or heat. The instability of MLD-deposited hybrid materials has limited their applications. In comparison, ALD has gained a much wider recognition in many areas, such as new energies,131–134 catalysis,135,136 and new materials.137,138 The last but not the least, MLD to date is restricted to a limited range of applications and has not received wide implementation. With the development of ideal MLD precursors and the improvement of resultant organic materials, we expect that the applications of MLD, in addition to those discussed in this review, will be greatly expanded into the areas of semiconductor,139 sensor,140 optics,141 biology,142 medical143,144 and smart materials.145

ACKNOWLEDGEMENTS Funding for this research was provided by the Center for Advanced Surface Engineering, under the National Science Foundation grant no. OIA-1457888 and the Arkansas EPSCoR Program, ASSET III.

REFERENCES 1. Mansoori, G. A.; Soelaiman, T. F., Nanotechnology—an introduction for the standards community. J. ASTM Int. 2005, 2 (6), 1–22. 2. James, S. L.; Adams, C. J.; Bolm, C.; Braga, D.; Collier, P.; Friščić, T.; Grepioni, F.; Harris, K. D. M.; Hyett, G.; Jones, W.; Krebs, A.; Mack, J.; Maini, L.; Orpen, A. G.;

Molecular Layer Deposition of Organic–Inorganic Hybrid Materials



63

Parkin, I. P.; Shearouse, W. C.; Steed, J. W.; Waddell, D. C., Mechanochemistry: Opportunities for new and cleaner synthesis. Chem. Soc. Rev. 2012, 41 (1), 413–447. 3. Takacs, L., The historical development of mechanochemistry. Chem. Soc. Rev. 2013, 42 (18), 7649–7659. 4. Dahl, J. A.; Maddux, B. L. S.; Hutchison, J. E., Toward Greener Nanosynthesis. Chem. Rev. 2007, 107 (6), 2228–2269. 5. Singh, J.; Wolfe, D. E., Review Nano and macro-structured component fabrication by electron beam-physical vapor deposition (EB-PVD). J. Mater. Sci. 2005, 40 (1), 1–26. 6. Cai, Z.; Liu, B.; Zou, X.; Cheng, H.-M., Chemical vapor deposition growth and applications of two-dimensional materials and their heterostructures. Chem. Rev. 2018, 118 (13), 6091–6133. 7. George, S. M., Atomic layer deposition: An overview. Chem. Rev. 2010, 110 (1), 111–131. 8. Meng, X., An overview of molecular layer deposition for organic and organic-inorganic hybrid materials: Mechanisms, growth characteristics, and promising applications. J. Mater. Chem. A 2017, 5 (35), 18326–18378. 9. George, S. M.; Lee, B. H.; Yoon, B.; Abdulagatov, A. I.; Hall, R. A., Metalcones: Hybrid organic–inorganic films fabricated using atomic and molecular layer deposition techniques. J. Nanosci. Nanotechnol. 2011, 11 (9), 7948–7955. 10. Lee, B. H.; Yoon, B.; Abdulagatov, A. I.; Hall, R. A.; George, S. M., Growth and properties of hybrid organic-inorganic metalcone films using molecular layer deposition techniques. Adv. Funct. Mater. 2013, 23 (5), 532–546. 11. Yoshimura, T.; Tatsuura, S.; Sotoyama, W., Polymer films formed with monolayer growth steps by molecular layer deposition. Appl. Phys. Lett. 1991, 59 (4), 482–484. 12. George, S. M.; Yoon, B.; Dameron, A. A., Surface chemistry for molecular layer deposition of organic and hybrid organic−inorganic polymers. Acc. Chem. Res. 2009, 42 (4), 498–508. 13. Zhou, W.; Leem, J.; Park, I.; Li, Y.; Jin, Z.; Min, Y.-S., Charge trapping behavior in organic-inorganic alloy films grown by molecular layer deposition from trimethylaluminum, p-phenylenediamine and water. J. Mater. Chem. 2012, 22 (45), 23935–23943. 14. Sundberg, P.; Karppinen, M., Organic and inorganic–organic thin film structures by molecular layer deposition: A review. Beilstein J. Nanotechnol. 2014, 5, 1104–1136. 15. Gregorczyk, K.; Knez, M., Hybrid nanomaterials through molecular and atomic layer deposition: Top down, bottom up, and in-between approaches to new materials. Prog. Mater Sci. 2016, 75, 1–37. 16. Yoshimura, T.; Tatsuura, S.; Sotoyama, W.; Matsuura, A.; Hayano, T., Quantum wire and dot formation by chemical vapor deposition and molecular layer deposition of onedimensional conjugated polymer. Appl. Phys. Lett. 1992, 60 (3), 268–270. 17. Yoshimura, T.; Oshima, A.; Kim, D.-I.; Morita, Y., Quantum dot formation in polymer wires by three-molecule molecular layer deposition (MLD) and applications to electrooptic/photovoltaic devices. ECS Trans. 2009, 25 (4), 15–25. 18. Yoshimura, T.; Ebihara, R.; Oshima, A., Polymer wires with quantum dots grown by molecular layer deposition of three source molecules for sensitized photovoltaics. J. Vac. Sci. Technol. A 2011, 29 (5), 051510. 19. Yoshimura, T.; Ishii, S., Effect of quantum dot length on the degree of electron localization in polymer wires grown by molecular layer deposition. J. Vac. Sci. Technol. A 2013, 31 (3), 031501. 20. Kim, A.; Filler, M. A.; Kim, S.; Bent, S. F., Layer-by-layer growth on Ge(100) via spontaneous urea coupling reactions. J. Am. Chem. Soc. 2005, 127 (16), 6123–6132. 21. Loscutoff, P. W.; Zhou, H.; Clendenning, S. B.; Bent, S. F., Formation of organic nanoscale laminates and blends by molecular layer deposition. ACS Nano 2010, 4 (1), 331–341. 22. Zhou, H.; Bent, S. F., Molecular layer deposition of functional thin films for advanced lithographic patterning. ACS Appl. Mater. Interfaces 2011, 3 (2), 505–511.

64

Optoelectronic Organic–Inorganic Semiconductor Heterojunctions

23. Zhou, H.; Toney, M. F.; Bent, S. F., Cross-linked ultrathin polyurea films via molecular layer deposition. Macromolecules 2013, 46 (14), 5638–5643. 24. Park, Y.-S.; Choi, S.-E.; Kim, H.; Lee, J. S., Fine-tunable absorption of uniformly aligned polyurea thin films for optical filters using sequentially self-limited molecular layer deposition. ACS Appl. Mater. Interfaces 2016, 8 (18), 11788–11795. 25. Shao, H.-I.; Umemoto, S.; Kikutani, T.; Okui, N., Layer-by-layer polycondensation of nylon 66 by alternating vapour deposition polymerization. Polymer 1997, 38 (2), 459–462. 26. Du, Y.; George, S. M., Molecular layer deposition of nylon 66 films examined using in situ FTIR spectroscopy. J. Phys. Chem. C 2007, 111 (24), 8509–8517. 27. Adamczyk, N. M.; Dameron, A. A.; George, S. M., Molecular layer deposition of poly(p-phenylene terephthalamide) films using terephthaloyl chloride and p-phenylenediamine. Langmuir 2008, 24 (5), 2081–2089. 28. Peng, Q.; Efimenko, K.; Genzer, J.; Parsons, G. N., Oligomer orientation in vapormolecular-layer-deposited alkyl-aromatic polyamide films. Langmuir 2012, 28 (28), 10464–10470. 29. Atanasov, S. E.; Losego, M. D.; Gong, B.; Sachet, E.; Maria, J.-P.; Williams, P. S.; Parsons, G. N., Highly conductive and conformal poly(3,4-ethylenedioxythiophene) (PEDOT) thin films via oxidative molecular layer deposition. Chem. Mater. 2014, 26 (11), 3471–3478. 30. Kim, D. H.; Atanasov, S. E.; Lemaire, P.; Lee, K.; Parsons, G. N., Platinum-free cathode for dye-sensitized solar cells using poly(3,4-ethylenedioxythiophene) (PEDOT) formed via oxidative molecular layer deposition. ACS Appl. Mater. Interfaces 2015, 7 (7), 3866–3870. 31. Miyamae, T.; Tsukagoshi, K.; Matsuoka, O.; Yamamoto, S.; Nozoye, H., Preparation of polyimide-polyamide random copolymer thin film by sequential vapor deposition polymerization. Jpn. J. Appl. Phys. 2002, 41 (2R), 746. 32. Loscutoff, P. W.; Lee, H.-B.-R.; Bent, S. F., Deposition of ultrathin polythiourea films by molecular layer deposition. Chem. Mater. 2010, 22 (19), 5563–5569. 33. Ivanova, T. V.; Maydannik, P. S.; Cameron, D. C., Molecular layer deposition of polyethylene terephthalate thin films. J. Vac. Sci. Technol. A 2012, 30 (1), 01A121. 34. Li, Y.-h.; Wang, D.; Buriak, J. M., Molecular layer deposition of thiol−ene multilayers on semiconductor surfaces. Langmuir 2010, 26 (2), 1232–1238. 35. Vasudevan, S. A.; Xu, Y.; Karwal, S.; van Ostaay, H. G. M. E.; Meesters, G. M. H.; Talebi, M.; Sudholter, E. J. R.; Ruud van Ommen, J., Controlled release from protein particles encapsulated by molecular layer deposition. Chem. Commun. 2015, 51 (63), 12540–12543. 36. Dameron, A. A.; Seghete, D.; Burton, B. B.; Davidson, S. D.; Cavanagh, A. S.; Bertrand, J. A.; George, S. M., Molecular layer deposition of alucone polymer films using trimethyl­aluminum and ethylene glycol. Chem. Mater. 2008, 20 (10), 3315–3326. 37. Park, Y.-S.; Kim, H.; Cho, B.; Lee, C.; Choi, S.-E.; Sung, M. M.; Lee, J. S., Intramolecular and Intermolecular Interactions in Hybrid Organic–Inorganic Alucone Films Grown by Molecular Layer Deposition. ACS Appl. Mater. Interfaces 2016, 8 (27), 17489–17498. 38. Choudhury, D.; Sarkar, S. K.; Mahuli, N., Molecular layer deposition of alucone films using trimethylaluminum and hydroquinone. J. Vac. Sci. Technol. A 2015, 33 (1), 01A115. 39. Bahlawane, N.; Arl, D.; Thomann, J.-S.; Adjeroud, N.; Menguelti, K.; Lenoble, D., Molecular layer deposition of amine-containing alucone thin films. Surf. Coat. Technol. 2013, 230, 101–105. 40. Miller, D. C.; Foster, R. R.; Jen, S.-H.; Bertrand, J. A.; Seghete, D.; Yoon, B.; Lee, Y.-C.; George, S. M.; Dunn, M. L., Thermomechanical properties of aluminum alkoxide (alucone) films created using molecular layer deposition. Acta Mater. 2009, 57 (17), 5083–5092.

Molecular Layer Deposition of Organic–Inorganic Hybrid Materials

65

41. Yoon, B.; Seghete, D.; Cavanagh, A. S.; George, S. M., Molecular layer deposition of hybrid organic−inorganic alucone polymer films using a three-step ABC reaction sequence. Chem. Mater. 2009, 21 (22), 5365–5374. 42. Gong, B.; Peng, Q.; Parsons, G. N., Conformal organic−inorganic hybrid network ­polymer thin films by molecular layer deposition using trimethylaluminum and g­ lycidol. J. Phys. Chem. B 2011, 115 (19), 5930–5938. 43. Lee, Y.; Yoon, B.; Cavanagh, A. S.; George, S. M., Molecular layer deposition of aluminum alkoxide polymer films using trimethylaluminum and glycidol. Langmuir 2011, 27 (24), 15155–15164. 44. Gong, B.; Parsons, G. N., Caprolactone ring-opening molecular layer deposition of organic-aluminum oxide polymer films. ECS J. Solid State Sci. Technol. 2012, 1 (4), P210-P215. 45. Abdulagatov, A. I.; Terauds, K. E.; Travis, J. J.; Cavanagh, A. S.; Raj, R.; George, S. M., Pyrolysis of titanicone molecular layer deposition films as precursors for conducting TiO2/carbon composite films. J. Phys. Chem. C 2013, 117 (34), 17442–17450. 46. Yoon, B.; O'Patchen, J. L.; Seghete, D.; Cavanagh, A. S.; George, S. M., Molecular layer deposition of hybrid organic-inorganic polymer films using diethylzinc and ethylene glycol. Chem. Vap. Deposition 2009, 15 (4–6), 112–121. 47. Peng, Q.; Gong, B.; VanGundy, R. M.; Parsons, G. N., “Zincone” zinc oxide−organic hybrid polymer thin films formed by molecular layer deposition. Chem. Mater. 2009, 21 (5), 820–830. 48. Yoon, B.; Lee, Y.; Derk, A.; Musgrave, C.; George, S., Molecular layer deposition of conductive hybrid organic-inorganic thin films using diethylzinc and hydroquinone. ECS Trans. 2011, 33 (27), 191–195. 49. Yoon, B.; Lee, B. H.; George, S. M., Molecular layer deposition of flexible, transparent and conductive hybrid organic-inorganic thin films. ECS Trans. 2011, 41 (2), 271–277. 50. Yoon, B.; Lee, B. H.; George, S. M., Highly conductive and transparent hybrid organic– inorganic zincone thin films using atomic and molecular layer deposition. J. Phys. Chem. C 2012, 116 (46), 24784–24791. 51. Han, K. S.; Sung, M. M., Molecular layer deposition of organic–inorganic hybrid films using diethylzinc and trihydroxybenzene. J. Nanosci. Nanotechnol. 2014, 14 (8), 6137–6142. 52. Cho, S.; Han, G.; Kim, K.; Sung, M. M., High-performance two-dimensional polydiacetylene with a hybrid inorganic–organic structure. Angew. Chem. Int. Ed. 2011, 50 (12), 2742–2746. 53. Brown, J. J.; Hall, R. A.; Kladitis, P. E.; George, S. M.; Bright, V. M., Molecular layer deposition on carbon nanotubes. ACS Nano 2013, 7 (9), 7812–7823. 54. Lee, B. H.; Anderson, V. R.; George, S. M., Molecular layer deposition of zircone and ZrO2/zircone alloy films: Growth and properties. Chem. Vap. Deposition 2013, 19 (4–6), 204–212. 55. Hall, R. A.; George, S. M.; Kim, Y.; Hwang, W.; Samberg, M. E.; Monteiro-Riviere, N. A.; Narayan, R. J., Growth of zircone on nanoporous alumina using molecular layer deposition. JOM 2014, 66 (4), 649–653. 56. Abdulagatov, A. I.; Hall, R. A.; Sutherland, J. L.; Lee, B. H.; Cavanagh, A. S.; George, S. M., Molecular layer deposition of titanicone films using TiCl4 and ethylene glycol or glycerol: Growth and properties. Chem. Mater. 2012, 24 (15), 2854–2863. 57. Van de Kerckhove, K.; Mattelaer, F.; Deduytsche, D.; Vereecken, P. M.; Dendooven, J.; Detavernier, C., Molecular layer deposition of "titanicone", a titanium-based hybrid material, as an electrode for lithium-ion batteries. Dalton Trans. 2016, 45 (3), 1176–1184. 58. Cao, Y.-Q.; Zhu, L.; Li, X.; Cao, Z.-Y.; Wu, D.; Li, A.-D., Growth characteristics of Ti-based fumaric acid hybrid thin films by molecular layer deposition. Dalton Trans. 2015, 44 (33), 14782–14792.

66

Optoelectronic Organic–Inorganic Semiconductor Heterojunctions

59. Chen, C.; Li, P.; Wang, G.; Yu, Y.; Duan, F.; Chen, C.; Song, W.; Qin, Y.; Knez, M., Nanoporous nitrogen-doped titanium dioxide with excellent photocatalytic activity under visible light irradiation produced by molecular layer deposition. Angew. Chem. Int. Ed. 2013, 52 (35), 9196–9200. 60. Lee, B. H.; Anderson, V. R.; George, S. M., Growth and properties of hafnicone and HfO2/hafnicone nanolaminate and alloy films using molecular layer deposition techniques. ACS Appl. Mater. Interfaces 2014, 6 (19), 16880–16887. 61. Van de Kerckhove, K.; Mattelaer, F.; Dendooven, J.; Detavernier, C., Molecular layer deposition of "vanadicone", a vanadium-based hybrid material, as an electrode for ­lithium-ion batteries. Dalton Trans. 2017, 46, 4542–4553. 62. Zhou, H.; Bent, S. F., Fabrication of organic interfacial layers by molecular layer deposition: Present status and future opportunities. J. Vac. Sci. Technol. A 2013, 31 (4), 040801. 63. Sarkar, D.; Ishchuk, S.; Taffa, D. H.; Kaynan, N.; Berke, B. A.; Bendikov, T.; Yerushalmi, R., Oxygen-deficient titania with adjustable band positions and defects; molecular layer deposition of hybrid organic–inorganic thin films as precursors for enhanced photocatalysis. J. Phys. Chem. C 2016, 120 (7), 3853–3862. 64. Ban, C.; George, S. M., Molecular layer deposition for surface modification of lithiumion battery electrodes. Adv. Mater. Interfaces 2016, 3, 1600762. 65. Gong, B.; Spagnola, J. C.; Parsons, G. N., Hydrophilic mechanical buffer layers and stable hydrophilic finishes on polydimethylsiloxane using combined sequential vapor infiltration and atomic/molecular layer deposition. J. Vac. Sci. Technol. A 2012, 30 (1), 01A156. 66. Groner, M. D.; Fabreguette, F. H.; Elam, J. W.; George, S. M., Low-temperature Al2O3 atomic layer deposition. Chem. Mater. 2004, 16 (4), 639–645. 67. Puurunen, R. L., Surface chemistry of atomic layer deposition: A case study for the trimethylaluminum/water process. J. Appl. Phys. 2005, 97 (12), 121301. 68. Vähä-Nissi, M.; Sievänen, J.; Salo, E.; Heikkilä, P.; Kenttä, E.; Johansson, L.-S.; Koskinen, J. T.; Harlin, A., Atomic and molecular layer deposition for surface modification. J. Solid State Chem. 2014, 214, 7–11. 69. Liang, X.; King, D. M.; Li, P.; George, S. M.; Weimer, A. W., Nanocoating hybrid polymer films on large quantities of cohesive nanoparticles by molecular layer deposition. AICHE J. 2009, 55 (4), 1030–1039. 70. Seghete, D.; Davidson, B. D.; Hall, R. A.; Chang, Y. J.; Bright, V. M.; George, S. M., Sacrificial layers for air gaps in NEMS using alucone molecular layer deposition. Sens. Actuators A Phys. 2009, 155 (1), 8–15. 71. Liang, X.; Yu, M.; Li, J.; Jiang, Y.-B.; Weimer, A. W., Ultra-thin microporous-mesoporous metal oxide films prepared by molecular layer deposition (MLD). Chem. Commun. 2009, (46), 7140–7142. 72. Qin, Y.; Yang, Y.; Scholz, R.; Pippel, E.; Lu, X.; Knez, M., Unexpected oxidation behavior of Cu nanoparticles embedded in porous alumina films produced by molecular layer deposition. Nano Lett. 2011, 11 (6), 2503–2509. 73. Lee, B. H.; Yoon, B.; Anderson, V. R.; George, S. M., Alucone Alloys with Tunable Properties Using Alucone Molecular Layer Deposition and Al2O3 Atomic Layer Deposition. J. Phys. Chem. C 2012, 116 (5), 3250–3257. 74. Jen, S.-H.; Lee, B. H.; George, S. M.; McLean, R. S.; Carcia, P. F., Critical tensile strain and water vapor transmission rate for nanolaminate films grown using Al2O3 atomic layer deposition and alucone molecular layer deposition. Appl. Phys. Lett. 2012, 101 (23), 234103. 75. Loebl, A. J.; Oldham, C. J.; Devine, C. K.; Gong, B.; Atanasov, S. E.; Parsons, G. N.; Fedkiw, P. S., Solid Electrolyte Interphase on Lithium-Ion Carbon Nanofiber Electrodes by Atomic and Molecular Layer Deposition. J. Electrochem. Soc. 2013, 160 (11), A1971-A1978.

Molecular Layer Deposition of Organic–Inorganic Hybrid Materials

67

76. Li, X.; Lushington, A.; Liu, J.; Li, R.; Sun, X., Superior stable sulfur cathodes of Li-S batteries enabled by molecular layer deposition. Chem. Commun. 2014, 50 (68), 9757–9760. 77. Lushington, A.; Liu, J.; Bannis, M. N.; Xiao, B.; Lawes, S.; Li, R.; Sun, X., A novel approach in controlling the conductivity of thin films using molecular layer deposition. Appl. Surf. Sci. 2015, 357, Part B, 1319–1324. 78. DuMont, J. W.; George, S. M., Pyrolysis of alucone molecular layer deposition films studied using in situ transmission fourier transform infrared spectroscopy. J. Phys. Chem. C 2015, 119 (26), 14603–14612. 79. Lemaire, P. C.; Oldham, C. J.; Parsons, G. N., Rapid visible color change and physical swelling during water exposure in triethanolamine-metalcone films formed by molecular layer deposition. J. Vac. Sci. Technol. A 2016, 34 (1), 01A134. 80. Oldham, C. J.; Gong, B.; Spagnola, J. C.; Jur, J. S.; Senecal, K. J.; Godfrey, T. A.; Parsons, G. N., Encapsulation and chemical resistance of electrospun nylon nanofibers coated using integrated atomic and molecular layer deposition. J. Electrochem. Soc. 2011, 158 (9), D549-D556. 81. Piper, D. M.; Travis, J. J.; Young, M.; Son, S.-B.; Kim, S. C.; Oh, K. H.; George, S. M.; Ban, C.; Lee, S.-H., Reversible high-capacity Si nanocomposite anodes for lithiumion batteries enabled by molecular layer deposition. Adv. Mater. 2014, 26 (10), 1596–1601. 82. Seghete, D.; Hall, R. A.; Yoon, B.; George, S. M., Importance of trimethylaluminum diffusion in three-step ABC molecular layer deposition using trimethylaluminum, ethanolamine, and maleic anhydride. Langmuir 2010, 26 (24), 19045–19051. 83. Liang, X.; Evanko, B. W.; Izar, A.; King, D. M.; Jiang, Y.-B.; Weimer, A. W., Ultrathin highly porous alumina films prepared by alucone ABC molecular layer deposition (MLD). Microporous Mesoporous Mater. 2013, 168, 178–182. 84. Gould, T. D.; Izar, A.; Weimer, A. W.; Falconer, J. L.; Medlin, J. W., Stabilizing Ni catalysts by molecular layer deposition for harsh, dry reforming conditions. ACS Catal. 2014, 4 (8), 2714–2717. 85. Huang, J.; Lucero, A. T.; Cheng, L.; Hwang, H. J.; Ha, M.-W.; Kim, J., HydroquinoneZnO nano-laminate deposited by molecular-atomic layer deposition. Appl. Phys. Lett. 2015, 106 (12), 123101. 86. Huang, J.; Zhang, H.; Lucero, A.; Cheng, L.; Kc, S.; Wang, J.; Hsu, J.; Cho, K.; Kim, J., Organic-inorganic hybrid semiconductor thin films deposited using molecular-atomic layer deposition (MALD). J. Mater. Chem. C 2016, 4 (12), 2382–2389. 87. Nilsen, O.; Haug, K. R.; Finstad, T.; Fjellvåg, H., Molecular hybrid structures by atomic layer deposition – Deposition of Alq3, Znq2 and Tiq4 (q = 8-hydroxyquinoline). Chem. Vap. Deposition 2013, 19 (4–6), 174–179. 88. Räupke, A.; Albrecht, F.; Maibach, J.; Behrendt, A.; Polywka, A.; Heiderhoff, R.; Helzel, J.; Rabe, T.; Johannes, H.-H.; Kowalsky, W.; Mankel, E.; Mayer, T.; Görrn, P.; Riedl, T., Conformal and highly luminescent monolayers of Alq3 prepared by gas-phase molecular layer deposition. ACS Appl. Mater. Interfaces 2014, 6 (2), 1193–1199. 89. Salmi, L. D.; Heikkilä, M. J.; Puukilainen, E.; Sajavaara, T.; Grosso, D.; Ritala, M., Studies on atomic layer deposition of MOF-5 thin films. Microporous Mesoporous Mater. 2013, 182, 147–154. 90. Salmi, L. D.; Heikkilä, M. J.; Vehkamäki, M.; Puukilainen, E.; Ritala, M.; Sajavaara, T., Studies on atomic layer deposition of IRMOF-8 thin films. J. Vac. Sci. Technol. A 2015, 33 (1), 01A121. 91. Ahvenniemi, E.; Karppinen, M., Atomic/molecular layer deposition: A direct gas-phase route to crystalline metal-organic framework thin films. Chem. Commun. 2016, 52 (6), 1139–1142.

68

Optoelectronic Organic–Inorganic Semiconductor Heterojunctions

92. Nisula, M.; Karppinen, M., Atomic/molecular layer deposition of lithium terephthalate thin films as high rate capability Li-ion battery anodes. Nano Lett. 2016, 16 (2), 1276–1281. 93. Kao, C.-Y.; Yoo, J.-W.; Min, Y.; Epstein, A. J., Molecular layer deposition of an organic-based magnetic semiconducting laminate. ACS Appl. Mater. Interfaces 2012, 4 (1), 137–141. 94. Kao, C.-Y.; Li, B.; Lu, Y.; Yoo, J.-W.; Epstein, A. J., Thin films of organic-based magnetic materials of vanadium and cobalt tetracyanoethylene by molecular layer deposition. J. Mater. Chem. C 2014, 2 (30), 6171–6176. 95. Zhang, B.; Chen, Y.; Li, J.; Pippel, E.; Yang, H.; Gao, Z.; Qin, Y., High efficiency Cu-ZnO hydrogenation catalyst: The tailoring of Cu-ZnO interface sites by molecular layer deposition. ACS Catal. 2015, 5 (9), 5567–5573. 96. Tynell, T.; Yamauchi, H.; Karppinen, M., Hybrid inorganic–organic superlattice structures with atomic layer deposition/molecular layer deposition. J. Vac. Sci. Technol. A 2014, 32 (1), 01A105. 97. Tynell, T.; Karppinen, M., ZnO: Hydroquinone superlattice structures fabricated by atomic/molecular layer deposition. Thin Solid Films 2014, 551, 23–26. 98. Lee, B. H.; Ryu, M. K.; Choi, S.-Y.; Lee, K.-H.; Im, S.; Sung, M. M., Rapid vapor-phase fabrication of organic−inorganic hybrid superlattices with monolayer precision. J. Am. Chem. Soc. 2007, 129 (51), 16034–16041. 99. Cha, S. H.; Oh, M. S.; Lee, K. H.; Im, S.; Lee, B. H.; Sung, M. M., Electrically stable low voltage ZnO transistors with organic/inorganic nanohybrid dielectrics. Appl. Phys. Lett. 2008, 92 (2), 023506. 100. Lee, K. H.; Choi, J.-M.; Im, S.; Lee, B. H.; Im, K. K.; Sung, M. M.; Lee, S., Lowvoltage-driven pentacene thin-film transistor with an organic-inorganic nanohybrid dielectric. Appl. Phys. Lett. 2007, 91 (12), 123502. 101. Lee, B. H.; Lee, K. H.; Im, S.; Sung, M. M., Vapor-phase molecular layer deposition of self-assembled multilayers for organic thin-film transistor. J. Nanosci. Nanotechnol. 2009, 9 (12), 6962–6967. 102. Huang, J.; Lee, M.; Lucero, A.; Kim, J., Organic-inorganic hybrid nano-laminates fabricated by ozone-assisted molecular-atomic layer deposition. Chem. Vap. Deposition 2013, 19 (4–6), 142–148. 103. Huang, J.; Lee, M.; Lucero, A. T.; Cheng, L.; Ha, M.-W.; Kim, J., 7-Octenyltrichrolosilane/ trimethyaluminum hybrid dielectrics fabricated by molecular-atomic layer deposition on ZnO thin film transistors. Jpn. J. Appl. Phys. 2016, 55 (6S1), 06GK04. 104. Lee, B. H.; Im, K. K.; Lee, K. H.; Im, S.; Sung, M. M., Molecular layer deposition of ZrO2-based organic–inorganic nanohybrid thin films for organic thin film transistors. Thin Solid Films 2009, 517 (14), 4056–4060. 105. Hashemi, F. S. M.; Prasittichai, C.; Bent, S. F., A new resist for area selective atomic and molecular layer deposition on metal–dielectric patterns. J. Phys. Chem. C 2014, 118 (20), 10957–10962. 106. Prasittichai, C.; Pickrahn, K. L.; Minaye Hashemi, F. S.; Bergsman, D. S.; Bent, S. F., Improving area-selective molecular layer deposition by selective SAM removal. ACS Appl. Mater. Interfaces 2014, 6 (20), 17831–17836. 107. McMahon, C. N.; Alemany, L.; Callender, R. L.; Bott, S. G.; Barron, A. R., Reaction of Al(tBu)3 with ethylene glycol: Intermediates to aluminum alkoxide (alucone) preceramic polymers. Chem. Mater. 1999, 11 (11), 3181–3188. 108. Baek, G.; Lee, S.; Lee, J.-H.; Park, J.-S., Air-stable alucone thin films deposited by molecular layer deposition using a 4-mercaptophenol organic reactant. J. Vac. Sci. Technol. A 2020, 38 (2), 022411. 109. Hagen, D. J.; Mai, L.; Devi, A.; Sainio, J.; Karppinen, M., Atomic/molecular layer deposition of Cu–organic thin films. Dalton Trans. 2018, 47 (44), 15791–15800.

Molecular Layer Deposition of Organic–Inorganic Hybrid Materials

69

110. Nilsen, O.; Klepper, K.; Nielsen, H.; Fjellvaåg, H., Deposition of organic-inorganic hybrid materials by atomic layer deposition. ECS Trans. 2008, 16 (4), 3–14. 111. Giedraityte, Z.; Sainio, J.; Hagen, D.; Karppinen, M., Luminescent metal-nucleobase network thin films by atomic/molecular layer deposition. J. Phys. Chem. C 2017, 121 (32), 17538–17545. 112. Giedraityte, Z.; Lopez-Acevedo, O.; Espinosa Leal, L. A.; Pale, V.; Sainio, J.; Tripathi, T. S.; Karppinen, M., Three-dimensional uracil network with sodium as a linker. J. Phys. Chem. C 2016, 120 (46), 26342–26349. 113. Mai, L.; Giedraityte, Z.; Schmidt, M.; Rogalla, D.; Scholz, S.; Wieck, A. D.; Devi, A.; Karppinen, M., Atomic/molecular layer deposition of hybrid inorganic–organic thin films from erbium guanidinate precursor. J. Mater. Sci. 2017, 52 (11), 6216–6224. 114. Isaeva, V. I.; Kustov, L. M., The application of metal-organic frameworks in catalysis (Review). Pet. Chem. 2010, 50 (3), 167–180. 115. Liu, X.-W.; Sun, T.-J.; Hu, J.-L.; Wang, S.-D., Composites of metal-organic frameworks and carbon-based materials: Preparations, functionalities and applications. J. Mater. Chem. A 2016, 4 (10), 3584–3616. 116. Murray, L. J.; Dinca, M.; Long, J. R., Hydrogen storage in metal-organic frameworks. Chem. Soc. Rev. 2009, 38 (5), 1294–1314. 117. Suh, M. P.; Park, H. J.; Prasad, T. K.; Lim, D.-W., Hydrogen storage in metal–organic frameworks. Chem. Rev. 2012, 112 (2), 782–835. 118. Khan, N. A.; Hasan, Z.; Jhung, S. H., Adsorptive removal of hazardous materials using metal-organic frameworks (MOFs): A review. J. Hazard. Mater. 2013, 244–245, 444–456. 119. Sabouni, R.; Kazemian, H.; Rohani, S., Carbon dioxide capturing technologies: A review focusing on metal organic framework materials (MOFs). Environ. Sci. Pollut. Res. 2014, 21 (8), 5427–5449. 120. Allendorf, M. D.; Bauer, C. A.; Bhakta, R. K.; Houk, R. J. T., Luminescent metalorganic frameworks. Chem. Soc. Rev. 2009, 38 (5), 1330–1352. 121. Horcajada, P.; Chalati, T.; Serre, C.; Gillet, B.; Sebrie, C.; Baati, T.; Eubank, J. F.; Heurtaux, D.; Clayette, P.; Kreuz, C.; Chang, J.-S.; Hwang, Y. K.; Marsaud, V.; Bories, P.-N.; Cynober, L.; Gil, S.; Ferey, G.; Couvreur, P.; Gref, R., Porous metal-organicframework nanoscale carriers as a potential platform for drug delivery and imaging. Nat Mater 2010, 9 (2), 172–178. 122. Vallet-Regí, M.; Balas, F.; Arcos, D., Mesoporous materials for drug delivery. Angew. Chem. Int. Ed. 2007, 46 (40), 7548–7558. 123. Wang, L.; Han, Y.; Feng, X.; Zhou, J.; Qi, P.; Wang, B., Metal–organic frameworks for energy storage: Batteries and supercapacitors. Coord. Chem. Rev. 2016, 307, Part 2, 361–381. 124. Xia, W.; Mahmood, A.; Zou, R.; Xu, Q., Metal-organic frameworks and their derived nanostructures for electrochemical energy storage and conversion. Energy Environ. Sci. 2015, 8 (7), 1837–1866. 125. Mahmood, A.; Guo, W.; Tabassum, H.; Zou, R., Metal-organic framework-based nanomaterials for electrocatalysis. Adv. Energy Mater. 2016, 6 (17), 1600423. 126. Gkaniatsou, E.; Sicard, C.; Ricoux, R.; Mahy, J.-P.; Steunou, N.; Serre, C., Metalorganic frameworks: A novel host platform for enzymatic catalysis and detection. Mater. Horiz. 2017, 4, 55–63. 127. Ahvenniemi, E.; Karppinen, M., In Situ atomic/molecular layer-by-layer deposition of inorganic–organic coordination network thin films from gaseous precursors. Chem. Mater. 2016, 28 (17), 6260–6265. 128. Penttinen, J.; Nisula, M.; Karppinen, M., Atomic/molecular layer deposition of s-block metal carboxylate coordination network thin films. Chem. –Eur. J. 2017, 23 (72), 18225–18231.

70

Optoelectronic Organic–Inorganic Semiconductor Heterojunctions

129. Olson, C. S.; Gangopadhyay, S.; Hoang, K.; Alema, F.; Kilina, S.; Pokhodnya, K., Magnetic exchange in MnII[TCNE] (TCNE = tetracyanoethylene) molecule-based magnets with two- and three-dimensional magnetic networks. J. Phys. Chem. C 2015, 119 (44), 25036–25046. 130. Miller, J. S., Magnetically ordered molecule-based materials. Chem. Soc. Rev. 2011, 40 (6), 3266–3296. 131. Meng, X.; Wang, X.; Geng, D.; Ozgit-Akgun, C.; Schneider, N.; Elam, J. W., Atomic layer deposition for nanomaterial synthesis and functionalization in energy technology. Mater. Horiz. 2017, 4 (2), 133–154. 132. Meng, X.; Yang, X. Q.; Sun, X. L., Emerging applications of atomic layer deposition for lithium-ion battery studies. Adv. Mater. 2012, 24 (27), 3589–3615. 133. Sun, Q.; Lau, K. C.; Geng, D.; Meng, X., Atomic and molecular layer deposition for superior lithium-sulfur batteries: Strategies, performance, and mechanisms. Batteries Supercaps 2018, 1 (2), 41–68. 134. Meng, X., Atomic-scale surface modifications and novel electrode designs for highperformance sodium-ion batteries via atomic layer deposition. J. Mater. Chem. A 2017, 5, 10127–10149. 135. O’Neill, B. J.; Jackson, D. H. K.; Lee, J.; Canlas, C.; Stair, P. C.; Marshall, C. L.; Elam, J. W.; Kuech, T. F.; Dumesic, J. A.; Huber, G. W., Catalyst design with atomic layer deposition. ACS Catal. 2015, 5 (3), 1804–1825. 136. Rimoldi, M.; Bernales, V.; Borycz, J.; Vjunov, A.; Gallington, L. C.; Platero-Prats, A. E.; Kim, I. S.; Fulton, J. L.; Martinson, A. B. F.; Lercher, J. A.; Chapman, K. W.; Cramer, C. J.; Gagliardi, L.; Hupp, J. T.; Farha, O. K., Atomic layer deposition in a metal–organic framework: Synthesis, characterization, and performance of a solid acid. Chem. Mater. 2017, 29 (3), 1058–1068. 137. Cai, J.; Han, X.; Wang, X.; Meng, X., Atomic layer deposition of two-dimensional layered materials: Processes, growth mechanisms, and characteristics. Matter 2020, 2, 587–630. 138. Zheng, X.; Lee, H.; Weisgraber, T. H.; Shusteff, M.; DeOtte, J.; Duoss, E. B.; Kuntz, J. D.; Biener, M. M.; Ge, Q.; Jackson, J. A.; Kucheyev, S. O.; Fang, N. X.; Spadaccini, C. M., Ultralight, ultrastiff mechanical metamaterials. Science 2014, 344 (6190), 1373–1377. 139. Lim, J. A.; Liu, F.; Ferdous, S.; Muthukumar, M.; Briseno, A. L., Polymer semiconductor crystals. Mater. Today 2010, 13 (5), 14–24. 140. Adhikari, B.; Majumdar, S., Polymers in sensor applications. Prog. Polym. Sci. 2004, 29 (7), 699–766. 141. Kara, P., Polymer optical fiber sensors—a review. Smart Mater. Struct. 2011, 20 (1), 013002. 142. Bajpai, A. K.; Bajpai, J.; Saini, R.; Gupta, R., Responsive polymers in biology and technology. Polym. Rev. 2011, 51 (1), 53–97. 143. Jagur-Grodzinski, J., Polymers for tissue engineering, medical devices, and regenerative medicine. Concise general review of recent studies. Polym. Adv. Technol. 2006, 17 (6), 395–418. 144. Maitz, M. F., Applications of synthetic polymers in clinical medicine. Biosurf. Biotribol. 2015, 1 (3), 161–176. 145. Jochum, F. D.; Theato, P., Temperature- and light-responsive smart polymer materials. Chem. Soc. Rev. 2013, 42 (17), 7468–7483.

4

Scanning Tunneling Microscope and Spectroscope on Organic–Inorganic Material Heterojunction Sadaf Bashir Khan and Shern Long Lee

CONTENTS 4.1 Introduction...................................................................................................... 71 4.2 Band Mapping; across a PN-Junction in a Nanorod........................................ 73 4.2.1 Nanorods and Junctions Characterization: Tunneling Current and Density of States������������������������������������������������������������� 76 4.2.2 Parallel PN-Junctions across Nanowires via One-Step Ex Situ Doping������������������������������������������������������������������� 78 4.3 Interfacial Band Mapping across Vertically Phased Separated Polymer/Fullerene Hybrid Solar Cells............................................ 79 4.4 Organic–Inorganic Hybrid Heterojunction...................................................... 86 4.4.1 Photocarrier Generations and Band Alignments at Perovskite/PbI2 Heterointerfaces�������������������������������������������������������� 86 4.4.2 Photocarrier Generations of Perovskites during Illumination............. 88 4.4.3 Band Alignments of Perovskites during Illumination.........................90 4.4.4 PbI2 Layer Thickness Dependence of ΔED......................................... 90 4.5 Outlook and Upcoming Challenges................................................................. 92 References................................................................................................................. 93

4.1 INTRODUCTION The operating mechanism of photovoltaic devices (PVs) comprises photogeneration of excitons, corresponding dissociation, and transference of holes and electrons to the opposite electrodes [1]. The effectiveness of these stages takes place in a sequence and can be influenced by active materials, exciton dissociation, which needs a junction in the form of a metal/organic or an organic/organic interface. The narrative of PVs symbolizes a junction between a stratum of donor and a stratum of acceptor molecules [2] such as a quantum dots (QDs) [QDs in a conjugated polymer matrix to form a hybrid bulk heterojunction (BHJ)] or BHJ [between electron-donor and 71

72

Optoelectronic Organic–Inorganic Semiconductor Heterojunctions

electron-acceptor molecules/polymers] [3,4]. Generally, heterojunction is defined as a contact of two materials having different electrical characteristics. These materials include polymers, elements, tiny molecules, or composites. Commonly, in a heterojunction, a donor is electron-rich than the acceptor. When the donor is excited via external stimuli such as light or electric pulse, it generates a negative charge in the acceptor, whereas the donor turns into a positive charge. The positive charges transfer via hole hopping and the negative charges via electron hopping, which takes place during charge transfer between molecules. Consequently, charges transport in the organic strata, which force charge to exchange in the electrodes. In hybrid BHJs, QDs, nanorods [range of II–VI, IV–VI], ternary or even quaternary semiconductors (SCs) have been integrated into appropriate polymer matrices [5–8]. Among nanostructures having different configurations, nanorods have an additional benefit as they offer one-dimensional (1D) conduction passageways for carriers in comparison to hopping transference via QDs. Among all junctions, exciton dissociation was confirmed via appropriate material selection, to generate a staggering bandgap or type-II band alignment between energy levels of the polymer-nanostructure at the interface [9]. Regarding this, scanning tunneling microscopy (STM) has made noteworthy contributions in comprehending the subprocesses underlying in various practical applications, specifically in photovoltaics. These procedures permit confined exploration of the ­materials’ electrical or optical characteristics. The spatially resolved measurements of surface photovoltage and photocurrent have been predominantly beneficial in comprehending the charge generation and separation in optoelectronic devices. In thin-film inorganic solar cells, charge separation does not take place at a heterojunction as expected, but instead, it occurs at a homojunction buried ∼50 nm within the absorbing layer [studied via Kelvin probe force microscopy (KPFM)] [10,11]. In organic photovoltaics, photocurrent has contributed to the understanding of the interplay concerning processing situations, blend morphology, configuration, or device enactment. This functional imaging technique discriminates STM from complementary structural characterization methods. The other techniques include SPM or atomic force microscopy (AFM). The AFM subdivided into different kinds, including photoconductive AFM, nearfield scanning optical microscopy, KPFM and time-resolved electric force microscopy. The band edges of low-dimensional nanostructures have been considered as chief influencing factors in designing an optoelectronic device. Different methods have been followed to monitor band edges such as doping, alloying or quantum confinement effect via lessening the length of active materials along with one to three directions [12–15]. The junctions generated between two semiconductors are in the form of type-I, [13,16] type-II/staggered [17–19], type-III or type-II broken gap [20–22]. The band-offset is characteristically thought-provoking system as far as the interface energy levels are concerned. Commonly, the band-offsets are engendered in coreshell nanoparticles, nanorods, nanowires, or in pn-junctions form in a nanorod [23– 28]. Henceforth, the interface offsets besides band edges are essential aspects of nanostructure engineering. STM so far remained an essential method in identifying the band edges of lower-dimensional coordination systems [29–33]. The capability to evaluate the nanostructures band edges has made the procedure distinctive in different aspects.

Scanning Tunneling Microscope and Spectroscope

73

The nanostructures extended from QDs toward intricate core-shells, hybrid coreshell arrangements in which the organic molecules generate a shell layer on inorganic nanocrystals [32,34]. In scientific exploration focused on practical applications, STM has been considered to observe the composite materials interface, that is, bulkheterojunction solar cells [13,35,36]. STM is a beneficial surface analytical technique that probes the local density of states (LDOS) having a high spatial and energetic resolution. In STM, a sharp metallic tip mechanically cut/chemically etched is brought into vicinity (∼0.3−1.0 nm) to a conducting or semiconducting surface. Generally, the tip is made of platinum–iridium (Pt/Ir) or tungsten (W) due to their chemical steadiness and mechanical resilience. An electrical bias is applied between the tip and the conducting substrate having ultrathin films. The electrons tunnel through the tip-sample gap, while skimming through the thin-film surface. The STM electronics display this current through a feedback loop to maintain a tipsample distance. The applied voltage polarity governs the direction of tunneling current. In the two bias directions, the tip might tunnel electrons to vacant states or accept electrons from filled states of the SC. Thus, the interconnection between the tip and the base electrode completes the circuit. A tunneling development happens when the magnitude of bias permits an energetic configuration between the tip work function and energy levels of the SC. At low temperature and low sample bias, the density of states (DOS) can be considered to be proportional to the tunneling conductance. The voltage dependence of tunneling current hence provides the location of the energy levels. In the differential conductance (dI/dV) spectrum, CB and VB edges appear as first peaks in the two bias directions. A dI/dV spectrum allows the derivation of the energy levels of the SC at the undergoing specified point of measurement. As it is a restricted localized mode of measurement, one may record many such spectra on various points of the SC monolayer. The acquired energies are plotted as histograms, which gives information regarding the location of CB and VB edges of the SC. One may deduce the DOS in the structure from the histograms distribution extending within the bandgap. As well, dI/dV versus voltage plots helps in energy-level mapping over the desired area. The nanoscale domains can be identified, and their corresponding local electronic characteristics can be probed simultaneously through dI/dV imaging. STM also helps in establishing a localized measurement model that applies to a single domain. In dI/dV images, the size of domains can be as small as a subnanometer due to tip shape. Hence, the material morphology under examination may depend on the substrate, information on the domains, substrate materials, and the externally applied parameters. The main objective of the present chapter is to deliver and demonstrate the materials science community with the competences, considerations, capabilities, and limitations linked with STM explorations focusing on optoelectronic organic–inorganic semiconductor heterojunction.

4.2 BAND MAPPING; ACROSS A PN-JUNCTION IN A NANOROD Earlier Amlan J. Pal and his colleagues investigate the band edges across generatejunction in a nanorod [37]. A single junction was established between Cu2S (p-type) and CdS (n-type) via controlled cationic exchange progression of CdS nanorods.

74

Optoelectronic Organic–Inorganic Semiconductor Heterojunctions

They illustrate nanorods comprising single materials and the distinct junction in a nanorod using an ultrahigh vacuum (UHV) scanning tunneling microscope (UHVSTM) at room temperature determining the conduction band (CB) and valence band (VB) edges at various points across the junction. They explore the band diagram of nanorod junctions to explore the salient structures of a diode, that is, p and n sections, depletion region, and band bending. The two distinctly designed nanostructures generate a junction grown via colloidal synthesis routes [38,39]. Between two distinct materials, the organic ligands act as stabilizers that perturb the interface or the junction itself [40]. The junction bands in a nanorod were mapped and visualize the electric field across the depletion region of the diode formed in a single nanorod through STM. The optical absorption spectra of CdS and Cu2S nanorods, and CdS|Cu2S junctions is presented in Figure 4.1a [i]. Two different reaction times

FIGURE 4.1  (a) Optical absorption spectra of CdS and Cu2S nanorods and CdS|Cu2S junctions, [i] In the nanorod junctions, (1) CdS|Cu2S (7 Min reaction time) and (2) CdS|Cu2S (10  Min) [for the cationic exchange process which controls the length of Cu2S section in nanorod junctions], [ii] STM topography of CdS, Cu2S nanorods, Cu2S|CdS, and [iii] Cu2S|CdS|Cu S nanorod junctions. Set-points for the approach of the tip were 0.4 nA at 2.0 V, which were used during STS measurements. (b) [i] STM topography of a CdS nanorod, [ii] displaying the points at which tunneling current versus voltage characteristics was noted, and [iii] DOS spectra of the I−V characteristics at the spots on the nanorod. The broken lines specify the conduction and VB edge’s location. (c) [i] STM topography of a CdS nanorod, [ii] displaying the points at which tunneling current versus voltage characteristics was noted, and [iii] DOS spectra of the I−V characteristics at the spots on the nanorod. The broken lines specify the conduction and VB edges location.

Scanning Tunneling Microscope and Spectroscope

75

FIGURE 4.1 (CONTINUED)  (d) [i] STM topography of a CdS|Cu2S nanorod junction showing the spots at which [ii] tunneling current versus voltage characteristics was recorded, [iii] DOS spectra of the I−V characteristics at the spots on the nanorod. The broken lines represent CB and VB edges; [iv] Schematic band diagrams of Cu2S and CdS nanorods before and after the formation of a pn-junction. (Reproduced with permission [37]).

(7  Min, 10 Min) were selected for the cationic exchange process in nanorod junctions, which impacts and control the length of Cu2S in the CdS|Cu2S junction. The optical absorption spectra demonstrate that toward short wavelength region, CdS shows absorption, however one of Cu2S shows absorption to the near-IR region. The absorbance in the 475–650 nm wavelengths appears due to Cu2S; due to the localized surface plasmon resonance arising because of copper deficiencies [38,41,42]. The CdS|Cu2S junction’s spectra were the sum of individual ones demonstrating the absorbance in the 475−650 nm regions upsurge with the growth in the length of Cu2S in CdS|Cu2S nanorod junctions.

76

Optoelectronic Organic–Inorganic Semiconductor Heterojunctions

4.2.1 Nanorods and Junctions Characterization: Tunneling Current and Density of States Figure 4.1a [ii] shows the STM topography of CdS, Cu2S nanorods and CdS|Cu2S junctions. The junctions encompassed Cu2S|CdS|Cu2S nanorods, Cu2S present at ends of CdS. The features of CdS and Cu2S were homogenous and uniform. The STM imaging was carried out in a constant-current mode. The Cu2S appears brighter than the CdS due to the higher conductivity of Cu2S, which enhances tip-to-nanorod distance in comparison to CdS. The variance in brightness and broadness leads to the classification of materials in a nanorod. The CdS and Cu2S nanorods diameter or their sections in a junction were nearly about 5 nm [43–47]. The probing of STM tip across the nanorods at different points helps in locating the conduction and VB edges. Therefore, the tunneling current versus applied voltage (I−V) was calculated via the STM tip. The tunneling current was noted after locating the STM tip overhead; the sample deactivating the scanning and feedback controls. Since the present structure looks like a double barrier tunnel junction (DBTJ) connecting the tip nanorod and nanorod substrate tunnel barriers [45,48]. The measurements were performed at various tip nanorod distances via different set-points. To evaluate the band edges, the DOS were estimated too. The bias was applied concerning the tip. That is why at the positive voltage region, the electrons can be introduced from the tip to the nanorod, symbolize the position of CBs. The I−V characteristics and STM topography at five different positions were recorded on CdS and Cu2S nanorods, as shown in Figure 4.1b and c [i, ii, iii]. In I−V characteristics, the asymmetry arises due to variation in the work function of the two electrodes. In the case of CdS nanorods, higher current in the positive voltage region is observed (Figure 4.1b) due to electrons insertion via tip into the CB of the n-type material, in comparison to the current in the negative voltage region. Correspondingly, for the p-type Cu2S, at the negative voltage region, the current remains higher due to facile hole injection to (i.e., electron extraction from the VB) (Figure 4.1c). The I−V characteristics across the nanorod nearly remain the same. It did not vary from one point to another, because single compositional material was being characterized along the nanorod. The DOS spectra of the nanorods demonstrated the location of conduction and VB edges in CdS and Cu2S. The positive and negative voltage at which peaks appear in DOS spectra signifies the conduction and VB’s location. The experimental result demonstrates that in CdS, the CB is positioned closer to the Fermi energy (fixed at 0 V). Likewise, in Cu2S, the VB is nearer to its Fermi energy. The DOS band, therefore, agrees with the n- and p-type characteristics of CdS and Cu2S nanorods. The position of band edges and DOS spectra intensity remain consistent throughout across nanorods. According to reported literature, the bandgaps of CdS and Cu2S are nearly 2.8 and 1.3 eV, respectively. The DOS depicts that Cu2S bandgap is lower than that of CdS. Besides this, the materials form a type-II band alignment when a junction formation takes place between CdS|Cu2S (single junction in a nanorod). The STM topography of CdS|Cu2S nanorod junctions is shown in Figure 4.1d [i]. The recorded I−V characteristics and corresponding DOS spectra (for determination of conduction and VB edges) on a single junction nanorod at various points having a smaller interval of distance are demonstrated in Figure 4.1d [ii, iii]. It is observed that the nanorod terminations resemble the individual material. On an individual point on the nanorod, at least 50 I−V characteristics were recorded to determine the

Scanning Tunneling Microscope and Spectroscope

77

conduction and VB edges from the DOS spectra. Figure 4.1d [iv] schematically represents the conduction and VB edges of a single pn-junction in a nanorod. The band edges at the terminals resemble the discrete nanorods, that is, CdS and Cu2S (Figure 4.1b and c). STM helps in mapping a single pn-junction signifying the conduction and VBs formed in a single nanorod having a length of about 70 nm. At the interface, band edges show the band bending along with the depletion region of a PN-junction diode. The outcomes allow one to envision the band locations along with the interface (between CdS and Cu2S sections in a pn-junction). The energy levels of the pn-­ junction at the interface coordinated well, like a conventional diode junction having a band bending of 0.3−0.4 eV. The experimental results hence demonstrate that the STM is a powerful technique in mapping across a pn-junction in a single nanorod the band edges via recording tunneling current at different points on the nanorods at room temperature. One can easily locate the DOS spectra, conduction and VB edges across the Cu2S and the CdS nanorods useful for heterojunctions optoelectronic devices. Accordingly, the mechanism of controlling and regulating the electronic, optical and magnetic characteristics of semiconductors (SCs) is correspondingly accomplished predominantly by the amalgamation of impurity atoms, that is, dopants, which modify the semiconductor’s electronic configuration and properties via creating surplus electrons (n-type) or holes (p-type). The systematic and accurate doping of SCs is the foundation of engendering photovoltaics, microelectronics, optoelectronic devices, sensors, or laser applications. Even though innovative, progressive approaches for the precise doping of fabricated SC assemblies have developed recently. Noteworthy progress has been made in this field, yet the SC doping at a nanoscale level still faces significant challenges in the semiconductor engineering [49–51]. Attaining highly oriented, organized, and ordered dopant distribution and in nanostructures is extremely important to enhance SCs optoelectrical performance. Recently, a pn-junction comprising the composite material was formed via introducing nanorods in a polymer matrix to form hybrid bulk heterojunction (BHJ) solar cells [52]. The pn-junctions were established in n-type CdS nanorods via an organized cationic exchange method. The p-type (Cu2S) generates at one end of CdS due to the selective reactivity of crystalline planes of the nanorods. In the nanorod, an epitaxial connection between Cu2S and CdS SC creates a depletion region forming a pn-junction, which is a classic illustration of type-II band alignment. The junction parted charge carriers during illumination via a drift of minority carriers across the depletion region. Thus, hybrid BHJs centered on Cu2S|CdS pn-junction nanorods in a polymer matrix represented as efficient solar cells as compared to similar BHJ devices with nanorods of individual materials (i.e., is Cu2S or CdS). The variation in the length of nanorods helps in controlling charge separation and carrier transport to enhance solar cell efficiency. The optimum proficiency of 3.7% at 1 sun intensity in BHJs was observed, comprising Cu2S|CdS (40:60, length-wise) pn-junctions in a P3HT matrix [52]. The efficiency is considerably developed in comparison to BHJ devices with nanorods of n-type CdS/p-type Cu2S. Besides this, optimized dopants (bismuth) insertion in lead sulfide (PbS) QDs establishes a BHJ solar cell having higher efficiency studied via STS generating a type-II BAND alignment [53]. Similar investigations on substrate solution were carried out via STM to explore the mechanism and examine electrical mapping or thermal stress, such as, analysis of the lateral and vertical phase separation of the photoactive layer in organic solar cells

78

Optoelectronic Organic–Inorganic Semiconductor Heterojunctions

comprising the solution-based polymer (donor) and fullerene (acceptor) in BHJ [54]. Amalgamation of an imidazolium-substituted polythiophene interlayer increased BJH efficiency significantly due to additional incorporated electric field promoting charge transfer [55]. In short, STM is appropriate to analyze semiconductors, metals, or biological samples competently [56].

4.2.2 Parallel PN-Junctions across Nanowires via One-Step Ex Situ Doping Earlier, a single-step technique was employed for the transformation of undoped silicon NWs (SiNWs) into nanoscale building blocks introducing parallel pn-junctions having homogenous and consistent dopant distributions [57]. The specified

FIGURE 4.2  (a) Schematic of [i] parallel pn-junction configuration formation across oriented NWs. [ii] Intrinsic Si NWs are transferred to a pretreated substrate with a monolayer containing boron functionalities and then covered by a second pretreated substrate with a phosphorus-containing annealed monolayer resulting in the controlled decomposition of the monolayers and dopant diffusion onto the NWs. (b) [i] STM/S spectroscopic dI/dV image recorded at −3.5 V sample bias, [ii] dopant concentration distribution across the NW junction, [iii] typical cross-sectional STM image of the NW obtained at a sample bias of 3.0 V. Inset: SEM secondary electron image displaying the geometrical structure of the NW from the crosssectional view, [iv] electrostatic potential distribution mapping, obtained by analyzing the STM/S data that demonstrates the formation of a junction in the NW, [v] CB edge and VB edge switched from p- to n-type across the NW, corresponding to the B- and P-doped poles, respectively, along the dashed arrow trajectory in (iv). The white dashed oval curve outlines the NW boundary. (Reproduced with permission [57])

Scanning Tunneling Microscope and Spectroscope

79

alignment of the resultant asymmetric configuration takes place comparative to the macroscopic framework. A parallel pn-junction formation takes place via a one-step process that is applied to SiNWs (Figure 4.2a [i & ii]). The doping method depends on a monolayer contact doping (MLCD) process, which develops monolayers as a source for dopant atoms. The MLCD method was earlier established to distribute extremely well-ordered development of doping elements on Si. The dopant concentration varies from several orders of magnitude from the modified interface into the bulk. The analogous pn-junction was measured via STM to acquire the electronic spatial characteristics of the junction generated across the NWs to reveal the association between the nanoscale topography and local electronic configuration via the cross-sectional profile of the NW at 100 K. The topographic STM imaging was accompanied at a continuous current of 0.4 nA having sample voltage of −0.3 V. The spectra were attained by operating STM in the current imaging tunneling spectroscopy (CITS) mode, various tunneling current descriptions were acquired at different sample bias voltages, Vs (4.0 to −4.0 V). The STM images and cross-sectional analyses of a SiNW are presented in Figure 4.2b [i,ii,iii,iv]. Figure 4.2a shows the formation of parallel pn-junctions across the SiNW by the one-step doping process. The Si-doped with B and P engenders a p- and n-type electronic characteristics, as demonstrated by the shift in the CB edge relative to the Fermi level (Figure 4.2b [v]). The two-dimensional (2D) cross-sectional spatial reliance of the DOS, at −3.5 V, for the parallel pn-junction configuration doped Si NW is presented in Figure 4.2b [i]. The STM/S spectroscopic data deliver dopant distributions quantitative mapping, electrostatic potential, and local band structures within the doped NWs (Figure 4.2b [ii, iv, v]). At the P-doped pole, the dopant surface concentration was 2.6 × 1019 cm3, whereas that at the B-doped pole was 1 × 1020 cm3. The doping concentration at the NW was 3 orders of reduced magnitude, demonstrating a compensated area approximately 1017 cm−3 n- and p-type dopant concentrations at NW core regions. Generally, the one-step NW doping procedure generated parallel n−i-p+ junctions through the NWs. Specifically, the n-type dopant dispersal displayed a diffusion profile with a leading dopant concentration at the NW interface interaction region, having a monotonic decline in dopant from the NW exterior, including core, which shows consistency. The post-growth and doping method proposes to tailor surface chemistry and control over dopant dispersal generating asymmetric configuration alignment that is important for employing nanoscale building blocks useful in electronic applications.

4.3 INTERFACIAL BAND MAPPING ACROSS VERTICALLY PHASED SEPARATED POLYMER/FULLERENE HYBRID SOLAR CELLS Above and beyond this, STM also helps in investigating the BHJ polymer solar cells. Usually, polymer solar cells gathered significant consideration in the fabrication of low-cost and mechanically flexible PVs as they support solution processing and patterning on flexible supports. The intensively explored and investigated materials for BHJ polymer solar cells comprise poly(3-hexylthiophene) (P3HT) and fullerene derivative phenyl-C61-butyric acid methyl ester (PCBM) blends, having power conversion of 4−5% [58–60]. In recent times high photovoltaic efficiencies (7−8%) have

80

Optoelectronic Organic–Inorganic Semiconductor Heterojunctions

been acquired via integrating s semiconducting polymers having small bandgap [6]. In semiconducting polymer electron acceptors generally, amalgam with polymers to form BHJs with a nanoscale interpenetrating donor/acceptor linkage due to short exciton diffusion length ( 2” effect in performance, which can form heterojunctions at their interfaces to generate new unique properties, making it a research hotspot. In this chapter, several combinations and applications of organic and inorganic semiconductor nanomaterials are described, as is referred to in Table 5.1. There are many ways and methods of combining inorganic semiconductor materials and organic semiconductor materials. This chapter provides a better description of the choice of inorganic and organic materials to build organic–inorganic semiconductor heterojunctions.

Heterojunction

n-type semiconductor

TPP-doped PFBT P3HT P3HT P3HT P3HT CuPc PDs BE P3HT LC polymer PPV PFO-DBT P3HT P3HT P3HT P3HT PANI spiro-MeOTAD PFH PVP PEDOT:PSS Tc CuPc MEH-PPV/ P3HT MEH-PPV PDINH g-C3N4 TNZnPc BBT P3HT PEDOT

CdS CdSe-tBT CdSe CdSe/PbS CdSe CdSe QDs CdS CdS CdS CdS ZnO ZnO ZnO ZnO ZnO ZnO ZnO ZnO ZnO ZnO ZnO ZnO CuZn2AlS4 ZnSe TiO2 TiO2 TiO2 TiO2 TiO2 TiO2

Application Photoelectrochemical Solar cell Solar cell Solar cell Photocathodefabrication Solar cell Photoelectric sensor Photocatalyst Solar cell Optoelectronic device Photodetector Solar cell Solar cell Photodetector Diode HCHO gas sensor Solar cell Diode/blue light detector. Photodetector Diode Photodetector Photovoltaic application Optoelectronic device Solar cell LEDs Photocatalyst Photocatalyst Photocatalyst Photocatalyst Photovoltaic application Photovoltaic application

Reference [19] [20] [21–24] [25] [26] [27] [28] [29] [30] [31] [32] [33] [34] [35] [36] [37] [38] [39] [40,41] [42] [43] [44] [45] [46] [47] [48] [49] [50] [51] [52] [53] (Continued)

117

Pdots/CdS P3HT/CdSe-tBT P3HT/CdSe P3HT/CdSe/PbS P3HT/CdSe CdSe/CuPc PDs/QDs BE-CdS P3HT/CdS LC/CdS CdS/PPV PFO-DBT/ZnO P3HT/ZnO P3HT/ZnO P3HT/ZnO P3HT/ZnO PANI/ZnO ZnO/spiro-MeOTAD PFH/ZnO PVP/ZnO PEDOT:PSS/ZnO ZnO/Tc ZnO/CuPc MEH-PPV/P3HT/CuZn2AlS4 ZnSe/MEH-PPV PDINH/TiO2 TiO2@g-C3N4 TNZnPc/TiO2 BBT/TiO2 TiO2/P3HT PEDOT/TiO2

p-type semiconductor

Organic–Inorganic Semiconducting Nanomaterial Heterojunctions

TABLE 5.1 Summary of different combinations of heterojunctions with their applications

Heterojunction

NPB CSA-PANI PANI PTCDA g-C3N4 g-C3N4 g-C3N4 C-337 Si PEDOT: PSS PPy TiO2 TiO2 TiO2 TiO2 PCBM PCBM DNTT SnO2 Se GCN g-C3N4 PANI Ag-Ag2O PPy g-C3N4 PTB7-Th PAE-1 PSBTBT PANI PDVT-10 PPy: Ru4POM

n-type semiconductor TiO2 F-Cl-TiO2 TiO2/SnO2 TiOPc N-LaTiO3 NiTi CaTiO3 n-Si SO Si PSi MAPbI3 CH3NH3PbI3 MASnI3 a-MAMnI3 MAPbI3 CH3NH3PbI3 FAPbI3 MASnI3 CH3NH3PbCl3 Ag3PO4 AgCl Ag2MoO4 TiO2@PPy Bi2WO6 BiOI BiI3 PbS PbS SnO2 MoS2 WO3

Application UV detector Photodetection Gas microsensor Photovoltaic cell Photocatalyst Photocatalyst Photocatalyst Solar cell Photovoltaic device Solar cell Electrode Hydrogen production Solar cell Solar cell Photodetector Phototransistor Solar cell Photodetector Sensor Photodetector Photocatalyst Photocatalyst Photocatalyst Photocatalyst Photoelectric detection Photocatalyst Solar cell Optical switch Solar cell Solar cell Optoelectronic application Photoanode

Reference [54] [55] [56] [57] [58] [59] [60] [61] [62] [63,64] [65] [66] [67] [68] [69] [70] [71] [72] [73] [74] [76] [77] [78] [79] [80] [81] [82] [84] [85] [86] [87] [88]

Optoelectronic Organic–Inorganic Semiconductor Heterojunctions

NPB/TiO2 CSA-PANI/F-Cl-TiO2 PANI/TiO2/SnO2 PTCDA/TiOPc g-C3N4/N-LaTiO3 g-C3N4/NiTi g-C3N4/CaTiO3 C-337/n-Si n-SO/P-Si PEDOT: PSS/Si PSi/PPy MAPbI3-TiO2 TiO2/CH3NH3PbI3 MASnI3/TiO2 a-MAMnI3/TiO2 MAPbI3/PCBM PCBM/CH3NH3PbI3 FAPbI3/DNTT MASnI3/SnO2 CH3NH3PbCl3/Se GCN–Ag3PO4 g-C3N4/AgCl PANI/Ag2MoO4 Ag/TiO2@PPy PPy @Bi2WO6 g-C3N4/BiOI BiI3/polymer PbS/PAE-1 PSBTBT/PbS PANI/SnO2 PDVT-10/MoS2 WO3/PPy

p-type semiconductor

118

TABLE 5.1 (Continued) Summary of different combinations of heterojunctions with their applications

Organic–Inorganic Semiconducting Nanomaterial Heterojunctions

119

ABBREVIATIONS AND ACRONYMS a-MAMnI3 Pb-free and amorphous MAMnI3 BBT poly (benzothiadiazole) BPA bisphenol A C-337 (2,3,6,7-tetrahydro-11-oxo-1H,5H,11H-[1]-benzopyrano[6,7,8ij]quinolizine-10-carbonitrile) CK-MB creatine kinase-methylene blue CSA camphor sulfonic acid CuPc copper phthalocyanine DNTT dinaphtho[2, 3-b: 2′, 3′-f] thieno-[3, 2-b] thiophene EDT 1, 2-ethanedithiol EQE external quantum efficiency ETL electron transporting layer g-C3N4 graphite carbon nitride GQDs graphene quantum dots HOMO highest occupied molecular orbital ISC short-circuit current kobs the pseudo first-order rate constant LC liquid crystal LDH layered double hydroxide Li-TFSI Li-bis-(trifluoromethane-sulfonyl) LUMO lowest occupied molecular orbital M-TiO2 mesoporous TiO2 Mehcf metal hexacyanoferrates MEH-PPV Poly [2-methoxy-5-(2-ethylhexyloxy)-1, 4-phenylenevinylene] NPB N, N′-Bis-(1-naphthalenyl)-N, N′-bis-phenyl-(1, 1′-biphenyl)-4, 4′-diamine NWs nanowires PAE-1 poly [sodium 2-(2-ethynyl-4-methoxyphenoxy) acetate] PANI polyaniline PBTTT poly (2,5-bis(3-al-kylthiophen-2-yl)thieno[3,2-b]thiophenes) PCBM [6,6]-phenyl-C61-butyric acid methyl ester PCE Power conversion efficiency PCPDTBT poly[2,6-(4,4-bis(2-ethylhexyl)-4H-cyclopenta [2,1-b;3,4-b] dithiophene)-alt-4,7(2,1,3-benzothiadiazole)] PDINH perylene-3, 4, 9, 10-tetracarboxylic diimide PDs polymer dots PDVT-10 poly[2,5-bis(2-decyltetradecyl)pyrrolo[3,4-c]pyrrole-1,4(2H,5H) dione-alt-5,5′-di(thiophen-2-yl)-2,2′-(E)-2-(2-(thiophen-2-yl) vinyl)-thiophene] PEC photoelectronchemical PEDOT poly (3, 4-ethylene dioxythiophene) PFBT poly [(9,9-dioctylfluorenyl-2,7-diyl)-co-(1,4-benzo-{2,1′, 3}-thiadazole)] PFH poly (9,9-dihexylfluorene)

120

Optoelectronic Organic–Inorganic Semiconductor Heterojunctions

PFO-DBT  poly [2,7-(9, 9-dioctylfluorene)-alt-4, 7-bis (thiophen-2-yl) benzo-2, 1, 3-thiadiazole] P3HT poly (3-hexylthiophene) PPy polypyrrole PPV poly (p-phenylene vinylene) PSBTBT  poly [2,6-(4, 4'–bis (2-ethylhexyl) dithieno [3,2-b: 2′, 3′-d] silole)-alt-4, 7 (2,1,3-benzothiadiazole) Psi porous silicon PSS poly (styrene sulfonate) PTB7-Th poly[4,8-bis(5-(2-ethylhexyl)thiophen-2yl) benzo[1,2-b,4,5-b′] dithiophene-2,6-diyl-alt-(4-(2-ethylhexyl)-3-fluorothieno[3,4b] thiophene-)-2-carboxylate-2-6-diyl] PTCDA perylenetertracarboxylic dianhydride spiro-MeOTAD 2,2′,7,7′-tetrakis-(N, N-di-p-methoxyphenylamine)-9, 9′-spirobifluorene PVP polyvinyl pyrrolidone QDs quantum dots RRT reverse recovery transient Ru4POM tetraruthenium polyoxometalate SO 1,3,3–Trimethy-lindolino-β-naphthopyrylospiran tBT tert-butylthiol Tc tetracene TiOPc titanyl phthalocyanine TNZnPc 2,9,16,23-tetranitrophthalocyanine zinc TP triphenylene TPP tetraphenylporphyrin UV ultraviolet VOC open-circuit voltage

REFERENCES 1. Zhang, J., et al., Evolution of epitaxial semiconductor nanodots and nanowires from supersaturated wetting layers. Chemical Society Reviews, 2015. 44(1): p. 26–39. 2. Iveland, J., et al., Direct measurement of Auger electrons emitted from a semiconductor light-emitting diode under electrical injection: Identification of the dominant mechanism for efficiency droop. Physical Review Letters, 2013. 110(17): p. 177406. 3. Walker, P.M., et al., Exciton polaritons in semiconductor waveguides. Applied Physics Letters, 2013. 102(1). 4. Zimmerman, J., R. Parameswaran, and B.Z. Tian, Nanoscale semiconductor devices as new biomaterials. Biomaterials Science, 2014. 2(5): p. 619–626. 5. Yousefi, M., et al., New role for nonlinear dynamics and chaos in integrated semiconductor laser technology. Physical Review Letters, 2007. 98(4): p. 044101. 6. Avrutin, V., N. Izyumskaya, and H. Morkoç, Semiconductor solar cells: Recent progress in terrestrial applications. Superlattices and Microstructures, 2011. 49(4): p. 337–364. 7. Rim, Y.S., et al., Printable ultrathin metal oxide semiconductor-based conformal biosensors. ACS Nano, 2015. 9(12): p. 12174–12181. 8. Suresh, S., Semiconductor nanomaterials, methods and applications: A review. Nanoscience and Nanotechnology, 2013.

Organic–Inorganic Semiconducting Nanomaterial Heterojunctions

121

9. Xu, J., et al., Enhancing visible-light-induced photocatalytic activity by coupling with wide-band-gap semiconductor: A case study on Bi2WO6/TiO2. Applied Catalysis B: Environmental, 2012. 111–112: p. 126–132. 10. Li, M., et al., Organic-inorganic heterojunctions toward high-performance ambipolar field-effect transistor applications. Advanced Electronic Materials, 2018. 4(9). 11. Umeda, T., et al., Improvement of characteristics on polymer photovoltaic cells composed of conducting polymer - fullerene systems. Synthetic Metals, 2005. 152(1–3): p. 93–96. 12. Amrollahi Bioki, H. and M. Borhani Zarandi, Electrical properties of organic–inorganic semiconductor heterojunction. Indian Journal of Physics, 2012. 86(6): p. 439–441. 13. Chen, C.-H., Hybrid organic on inorganic semiconductor heterojunction. Journal of Materials Science Materials in Electronics, 2006. 17(12): p. 1047–1053. 14. Chen, Y., et al., Self-assembled organic-inorganic hybrid nanocomposite of a perylenetetracarboxylic diimide derivative and CdS. Langmuir, 2010. 26(15): p. 12473–12478. 15. Liu, R., Hybrid organic/inorganic nanocomposites for photovoltaic cells. Materials (Basel), 2014. 7(4): p. 2747–2771. 16. Mir, S.H., et al., Review—organic-inorganic hybrid functional materials: An integrated platform for applied technologies. Journal of The Electrochemical Society, 2018. 165(8): p. B3137–B3156. 17. Aleshin, A.N., Organic optoelectronics based on polymer – inorganic nanoparticle composite materials. Physics-Uspekhi, 2013. 56(6): p. 627–632. 18. Li, S., et al., Nanocomposites of polymer and inorganic nanoparticles for optical and magnetic applications. Nano Reviews, 2010. 1. 19. Mao, L., et al., Organic-inorganic nanoparticles molecularly imprinted photoelectrochemical sensor for α-Solanine based on p-type polymer dots and n-CdS heterojunction. Analytica Chimica Acta, 2019. 1059: p. 94–102. 20. Greaney, M.J., et al., Improving open circuit potential in hybrid P3HT:CdSe bulk heterojunction solar cells via colloidal tert-butylthiol ligand exchange. ACS Nano, 2012. 6(5): p. 4222–4230. 21. Lim, J., et al., Modular fabrication of hybrid bulk heterojunction solar cells based on breakwater-like CdSe tetrapod nanocrystal network infused with P3HT. The Journal of Physical Chemistry C, 2014. 118(8): p. 3942–3952. 22. Xu, W., et al., Efficient Organic/Inorganic Hybrid Solar Cell Integrating Polymer Nanowires and Inorganic Nanotetrapods. Nanoscale Research Letters, 2017. 12(1). 23. Zhou, Y., et al., Improved efficiency of hybrid solar cells based on non-ligand-exchanged CdSe quantum dots and poly(3-hexylthiophene). Applied Physics Letters, 2010. 96(1): p. 013304. 24. Khan, M.A., U. Farva, and Y.M. Kang, Pyrolytic synthesis of densely packed grown CdSe nanoparticles using a dual function octylamine for hybrid solar cells. Materials Letters, 2014. 123: p. 62–65. 25. Tan, F., et al., Core/shell-shaped CdSe/PbS nanotetrapods for efficient organic–inorganic hybrid solar cells. Journal of Materials Chemistry A, 2014. 2(35). 26. Lai, L.H., et al., Organic-inorganic hybrid solution-processed H(2)-evolving photocathodes. ACS Applied Materials & Interfaces, 2015. 7(34): p. 19083–19090. 27. Saha, S.K., A. Guchhait, and A.J. Pal, Organic/inorganic hybrid pn-junction between copper phthalocyanine and CdSe quantum dot layers as solar cells. Journal of Applied Physics, 2012. 112(4): p. 044507. 28. Wang, Q., et al., Semiconducting organic–inorganic nanodots heterojunctions: Platforms for general photoelectrochemical bioanalysis application. Analytical Chemistry, 2018. 90(6): p. 3759–3765. 29. Zhang, X.H., et al., Robust visible/near-infrared light driven hydrogen generation over Z-scheme conjugated polymer/CdS hybrid. Applied Catalysis B-Environmental, 2018. 224: p. 871–876.

122

Optoelectronic Organic–Inorganic Semiconductor Heterojunctions

30. Ren, S., et al., Inorganic–organic hybrid solar cell: Bridging quantum dots to conjugated polymer nanowires. Nano Letters, 2011. 11(9): p. 3998–4002. 31. Yuan, K., L. Chen, and Y. Chen, Direct anisotropic growth of CdS nanocrystals in thermotropic liquid crystal templates for heterojunction optoelectronics. Chemistry - A European Journal, 2014. 20(36): p. 11488–11495. 32. Lin, H.W., et al., Constructing a green light photodetector on inorganic/organic semiconductor homogeneous hybrid nanowire arrays with remarkably enhanced photoelectric response. ACS Applied Materials & Interfaces, 2019. 11(10): p. 10146–10152. 33. Sharma, R., et al., Role of zinc oxide and carbonaceous nanomaterials in non-fullerenebased polymer bulk heterojunction solar cells for improved cost-to-performance ratio. Journal of Materials Chemistry A, 2015. 3(44): p. 22227–22238. 34. Pei, J., et al., Influence of organic interface modification layer on the photoelectric properties of ZnO-based hybrid solar cells. Journal of Photochemistry and Photobiology A: Chemistry, 2018. 364: p. 551–557. 35. Ouyang, B.S., K.W. Zhang, and Y. Yang, Self-powered UV photodetector array based on P3HT/ZnO nanowire array heterojunction. Advanced Materials Technologies, 2017. 2(12). 36. Pandey, R.K., et al., Interface engineering for enhancement in performance of organic/ inorganic hybrid heterojunction diode. Organic Electronics, 2017. 45: p. 26–32. 37. Tai, H.L., et al., The enhanced formaldehyde-sensing properties of P3HT-ZnO hybrid thin film OTFT sensor and further insight into its stability. Sensors, 2015. 15(1): p. 2086–2103. 38. Tonpe, D.A., et al., Development of organic/inorganic PANI/ZnO 1D nanostructured hybrid thin film solar cell by soft chemical route. Journal of Materials Science-Materials in Electronics, 2019. 30(17): p. 16056–16064. 39. Xie, Y.R., et al., High performance blue light detector based on ZnO nanowire arrays. Applied Optics, 2019. 58(5): p. 1242–1245. 40. Li, X.Y., et al., A self-powered nano-photodetector based on PFH/ZnO nanorods organic/inorganic heterojunction. Results in Physics, 2018. 8: p. 468–472. 41. Song, J., et al., Poly(9,9-dihexylfluorene)/ZnO nanoparticles based inorganic/organic heterojunction structure: Electrical and photoconductivity properties. Journal of Nanoscience and Nanotechnology, 2016. 16(6): p. 6005–6010. 42. Mokhtari, H. and M. Benhaliliba, Organic-inorganic Au/PVP/ZnO/Si/Al semiconductor heterojunction characteristics. Journal of Semiconductors, 2017. 38(11). 43. Trung, T.Q., et al., An omnidirectionally stretchable photodetector based on organicinorganic heterojunctions. ACS Applied Materials & Interfaces, 2017. 9(41): p. 35958–35967. 44. Luka, G., et al., Electrical properties of zinc oxide – Tetracene heterostructures with different n-type ZnO films. Organic Electronics, 2017. 45: p. 240–246. 45. Padma, N., et al., Performance comparison of p-n junction diodes using zinc oxide and copper phthalocyanine hybrid nanocomposites and bilayer heterostructures. Nano, 2014. 9(6). 46. Chaudhary, D.K., et al., Bulk-heterojunction hybrid solar cells with non-toxic, earth abundant stannite phase CuZn2AlS4 nanocrystals. Thin Solid Films, 2018. 649: p. 202–209. 47. Chen, G.-H., S.-J. Ho, and H.-S. Chen, Cubic zincblende ZnSe nanowires with an entangling structure grown via oriented attachment and their application in organic–inorganic heterojunction light-emitting diodes. The Journal of Physical Chemistry C, 2014. 118(44): p. 25816–25822. 48. Li, X., et al., Self-assembled supramolecular system PDINH on TiO2 surface enhances hydrogen production. Journal of Colloid and Interface Science, 2018. 525: p. 136–142.

Organic–Inorganic Semiconducting Nanomaterial Heterojunctions

123

49. Xiao, G., et al., Visible-light-driven activity and synergistic mechanism of TiO2@gC3N4 heterostructured photocatalysts fabricated through a facile and green procedure for various toxic pollutants removal. Nanotechnology, 2018. 29(31). 50. Zhang, Q., et al., Tetranitrophthalocyanine Zinc/TiO2 nanofibers organic-inorganic heterostructures with enhanced visible photocatalytic activity. Nano, 2017. 12(10). 51. Hou, H.J., et al., Conjugated microporous poly(benzothiadiazole)/TiO2 heterojunction for visible-light-driven H-2 production and pollutant removal. Applied Catalysis B-Environmental, 2017. 203: p. 563–571. 52. Qin, Y., et al., Top-down strategy toward versatile graphene quantum dots for organic/ inorganic hybrid solar cells. ACS Sustainable Chemistry & Engineering, 2015. 3(4): p. 637–644. 53. Szkoda, M., et al., Fabrication and photoactivity of organic-inorganic systems based on titania nanotubes and PEDOT containing redox centres formed by different Prussian Blue analogues. Journal of Alloys and Compounds, 2017. 723: p. 498–504. 54. Zhang, D.Z., et al., Organics filled one-dimensional TiO2 nanowires array ultraviolet detector with enhanced photo-conductivity and dark-resistivity. Nanoscale, 2017. 9(26): p. 9095–9103. 55. Sharma, S., et al., Fluorine-chlorine co-doped TiO2/CSA doped polyaniline based high performance inorganic/organic hybrid heterostructure for UV photodetection applications. Sensors and Actuators a-Physical, 2017. 261: p. 94–102. 56. Subramanian, E., P. Santhanamari, and C. Murugan, Sensor functionality of conducting polyaniline-metal oxide (TiO2/SnO2) hybrid materials films toward benzene and toluene vapors at room temperature. Journal of Electronic Materials, 2018. 47(8): p. 4764–4771. 57. Jarka, P., et al., Manufacture of photovoltaic cells with hybrid organic-inorganic bulk heterojunction. Materials and Manufacturing Processes, 2018. 33(8): p. 912–922. 58. Rakibuddin, M., H. Kim, and M.E. Khan, Graphite-like carbon nitride (C3N4) modified N-doped LaTiO3 nanocomposite for higher visible light photocatalytic and photo-electrochemical performance. Applied Surface Science, 2018. 452: p. 400–412. 59. Yazdani, D., A.A. Zinatizadeh, and M. Joshaghani, Organic-inorganic Z-scheme g-C3N4-NiTi-layered double hydroxide films for photocatalytic applications in a fixedbed reactor. Journal of Industrial and Engineering Chemistry, 2018. 63: p. 65–72. 60. Kumar, A., et al., Perovskite-structured CaTiO3 coupled with g-C3N4 as a heterojunction photocatalyst for organic pollutant degradation. Beilstein Journal of Nanotechnology, 2018. 9: p. 671–685. 61. Nawar, A.M. and M.M. Makhlouf, Electrical and photovoltaic responses of an Au/ Coumarin 337/n-Si/Sb-Au hybrid organic-inorganic solar cell. Journal of Electronic Materials, 2019. 48(9): p. 5771–5784. 62. El-Nahass, M.M. and W.M. Desoky, Fabrication and electrical characterization of n-1,3,3-trimethylindolino-beta-naphthopyrylospiran/p-Si (organic/inorganic) heterojunction for application in photovoltaic device. Optik, 2018. 171: p. 44–50. 63. Mahato, S., et al., High efficiency ITO-free hybrid solar cell using highly conductive PEDOT:PSS with co-solvent and surfactant treatments. Materials Letters, 2017. 186: p. 165–167. 64. Zhang, J., S.T. Lee, and B.Q. Sun, Effect of series and shunt resistance on organicinorganic hybrid solar cells performance. Electrochimica Acta, 2014. 146: p. 845–849. 65. Harraz, F.A., Electrochemical formation of a novel porous silicon/polypyrrole hybrid structure with enhanced electrical and optical characteristics. Journal of Electroanalytical Chemistry, 2014. 729: p. 68–74. 66. Luo, J., et al., Organic-inorganic hybrid perovskite – TiO2 nanorod arrays for efficient and stable photoelectrochemical hydrogen evolution from HI splitting. Materials Today Chemistry, 2019. 12: p. 1–6.

124

Optoelectronic Organic–Inorganic Semiconductor Heterojunctions

67. Zhou, H.W., et al., Hole-conductor-free, metal-electrode-free TiO2/CH3NH3PbI3 heterojunction solar cells based on a low-temperature carbon electrode. Journal of Physical Chemistry Letters, 2014. 5(18): p. 3241–3246. 68. Zhang, W.N., et al., Lead-free organic-inorganic hybrid perovskite heterojunction composites for photocatalytic applications. Catalysis Science & Technology, 2017. 7(13): p. 2753–2762. 69. Zhang, X.X., et al., Lead-free and amorphous organic-inorganic hybrid materials for photovoltaic applications: mesoscopic CH3NH3MnI3/TiO2 heterojunction. RSC Advances, 2017. 7(59): p. 37419–37425. 70. Jeong, H. and J.K. Lee, Organic–inorganic hybrid ternary bulk heterojunction of nanostructured perovskite–low bandgap polymer–PCBM for improved efficiency of organic solar cells. ACS Applied Materials & Interfaces, 2015. 7(51): p. 28459–28465. 71. Rahaq, Y., et al., Highly reproducible perovskite solar cells via controlling the morphologies of the perovskite thin films by the solution-processed two-step method. Journal of Materials Science-Materials in Electronics, 2018. 29(19): p. 16426–16436. 72. Luo, L.B., et al., A highly sensitive perovskite/organic semiconductor heterojunction phototransistor and its device optimization utilizing the selective electron trapping effect. Advanced Optical Materials, 2019. 7(13). 73. Chen, Y., et al., Light enhanced room temperature resistive NO2 sensor based on a goldloaded organic-inorganic hybrid perovskite incorporating tin dioxide. Microchimica Acta, 2019. 186(1). 74. Chen, H.-Y., et al., Fabrication and photoelectric properties of organic-inorganic broadspectrum photodetectors based on Se microwire/perovskite heterojunction. Chinese Optics, 2019. 12(5): p. 1057–1063. 75. Huang, K., et al., One-step synthesis of Ag3PO4/Ag photocatalyst with visible-light photocatalytic activity. Materials Research, 2015. 18(5): p. 939–945. 76. Nagajyothi, P.C., et al., Photocatalytic dye degradation and hydrogen production activity of Ag3PO4/g-C3N4 nanocatalyst. Journal of Materials Science-Materials in Electronics, 2019. 30(16): p. 14890–14901. 77. Matheswaran, P., P. Thangavelu, and B. Palanivel, Carbon dot sensitized integrative g-C3N4/AgCl Hybrids: An synergetic interaction for enhanced visible light driven photocatalytic process. Advanced Powder Technology, 2019. 30(8): p. 1715–1723. 78. Mondal, P., et al., Facile fabrication of novel hetero-structured organic-inorganic highperformance nanocatalyst: A smart system for enhanced catalytic activity toward ciprofloxacin degradation and oxygen reduction. ACS Applied Nano Materials, 2018. 1(11): p. 6015–6026. 79. Kumar, R., R.M. El-Shishtawy, and M.A. Barakat, Synthesis and characterization of Ag-Ag2O/TiO2@polypyrrole heterojunction for enhanced photocatalytic degradation of methylene blue. Catalysts, 2016. 6(6). 80. Zhu, L.-B., et al., Enhanced organic−inorganic heterojunction of polypyrrole@Bi2WO6: Fabrication and application for sensitive photoelectrochemical immunoassay of creatine kinase-MB. Biosensors and Bioelectronics, 2019. 140: p. 111349. 81. Chang, C., et al., Novel mesoporous graphite carbon nitride/BiOI heterojunction for enhancing photocatalytic performance under visible-light irradiation. ACS Applied Materials & Interfaces, 2014. 6(7): p. 5083–5093. 82. Ma, S., et al., Vertically oriented BiI3 template featured BiI3/polymer heterojunction for high photocurrent and long-term stable solar cells. ACS Applied Materials & Interfaces, 2019. 11(35): p. 32509–32516. 83. Kumar, R., et al., Preparation of nanocrystalline Sb doped PbS thin films and their structural, optical, and electrical characterization. Superlattices and Microstructures, 2014. 75: p. 601–612.

Organic–Inorganic Semiconducting Nanomaterial Heterojunctions

125

84. Wurst, K.M., et al., Correlated, dual-beam optical gating in coupled organic-inorganic nanostructures. Angewandte Chemie, 2018. 130(36): p. 11733–11737. 85. Nam, M., et al., Broadband-absorbing hybrid solar cells with efficiency greater than 3% based on a bulk heterojunction of PbS quantum dots and a low-bandgap polymer. Journal of Materials Chemistry A, 2014. 2(11): p. 3978. 86. Murugan, C., E. Subramanian, and D.P. Padiyan, p-n Heterojunction formation in polyaniline-SnO2 organic-inorganic hybrid composite materials leading to enhancement in sensor functionality toward benzene and toluene vapors at room temperature. Synthetic Metals, 2014. 192: p. 106–112. 87. Yan, J., et al., Ambipolar charge transport in an organic/inorganic van der Waals p-n heterojunction. Journal of Materials Chemistry C, 2018. 6(47): p. 12976–12980. 88. Jeon, D., et al., WO3/conducting polymer heterojunction photoanodes for efficient and stable photoelectrochemical water splitting. ACS Applied Materials & Interfaces, 2018. 10(9): p. 8036–8044.

6

Organic–Inorganic Heterojunction Nanowires Yuan Yao and Yanbing Guo

CONTENTS 6.1 Introduction: Background and Driving Forces.............................................. 127 6.2 The Synthetic Methods of Organic–Inorganic Heterojunction Nanowires....... 128 6.2.1 Solution Phase Method...................................................................... 128 6.2.2 Template Method Combined with Electrochemical Polymerization...... 130 6.2.2.1 Template Method Combined with Pressure Injection���������������������������������������������������������������������������������� 132

6.2.3 Vapor–Liquid–Solid Method............................................................. 132 6.3 The Applications of Organic–Inorganic Heterojunction Nanowires............. 132 6.3.1 Field Emission................................................................................... 133 6.3.2 Diode Rectification............................................................................ 136 6.3.3 Solar Cells.......................................................................................... 138 6.3.4 Photoelectric Detection...................................................................... 140 6.3.5 Logic Gates........................................................................................ 140 6.4 Summary and Perspective.............................................................................. 143 References............................................................................................................... 143

6.1 INTRODUCTION: BACKGROUND AND DRIVING FORCES Among crystalline materials, inorganic crystals, such as silicon, were the fundamental building blocks of modern electronics and microelectronics. It was not surprising that the preparation and controlled growth of crystalline nanostructures with various morphologies and orientations was a hot research topic [1,2]. Compared with inorganic materials, organic materials have the advantages of low cost, easy molecular tailoring for property optimization, high flexibility, easy large-scale processing as well as compatibility with lightweight plastic substrates [3–5]. Thus, π-conjugated polymer- and small molecule-based functional organic nanomaterials are considered to be good candidates for the next generation miniature, flexible consumer electronic devices [6–10]. However, the short-range order, amorphous organic materials always share the problem of low-carrier transport ability and low-performance stability [11]. Organic and inorganic materials can be combined by covalent bonds or other interaction forces to form organic–inorganic heterojunction. This strategy not only preserves the original characteristics of each material but also forms new properties 127

128

Optoelectronic Organic–Inorganic Semiconductor Heterojunctions

through interface interaction, thus realizing the multifunctional materials. On the other hand, one-dimensional nanomaterials have attracted much attention due to their large length-diameter ratio and unique optical and photoelectric properties [12]. The devices made of one-dimensional materials have the characteristics of multifunction and can be used in field-emission diodes, solar cells, photoelectric detection, and other fields [13]. In these devices, the formation of heterojunction has a critical and significant effect on the material properties. Based on the above analysis, one-dimensional organic–inorganic heterojunction nanowires-based devices have multifunction, unique electrical, optical, and tunable characteristics, and therefore have potential development and application prospects in the fields of electronics, photoelectricity, and catalysis, etc. [11,14]. Strong interactions between organic and inorganic components could result in new or enhanced physical or chemical properties relative to that of a single component, thus achieving synergistic properties [14]. At present, one-dimensional organic–inorganic heterojunction nanowires have made important progress in both synthesis methodology and material properties. There are three research directions: (1) The preparation of novel organic–inorganic heterojunction by the reasonable design of molecular structure and the regulation of interfacial forces; (2) The studies of growth process and formation mechanism; (3)  The detection of improved physical or chemical properties. Therefore, we reviewed the recent progress of organic–inorganic heterojunction nanowires, including preparation methods and application fields. Besides, the problems that need to be solved urgently and the future development direction of this field were pointed out to better promote the continuous progress of this field.

6.2 THE SYNTHETIC METHODS OF ORGANIC–INORGANIC HETEROJUNCTION NANOWIRES Up to now, there are four methods to fabricate organic–inorganic heterojunction nanowires, including chemical vapor deposition, solution phase method and template method. Even though the growth mechanism of nanowires via these methods is different, the growth of nanowires (crystals) is affected by both kinetics and thermodynamics [13]. In this section, the specific synthesis paths of each method are introduced, and the growth mechanism of organic–inorganic heterojunction nanowires is emphasized. This provides theoretical guidance and reference for the design and development of the next generation of higher-quality organic–inorganic heterojunction nanowires and novel synthesis methods.

6.2.1 Solution Phase Method The solution phase method is the simplest method to synthesize organic–inorganic heterojunction nanowires: the substrate was immersed in the solution of organic precursors, and then it was taken out and cleaned to obtain organic–inorganic heterojunction nanowires after a period of reaction [15]. In 2009, Huibiao Liu et al. used zinc oxide (ZnO) with wide band gap to modify Cu-TCNQ. The ZnO-CuTCNQ hybrid NWs was prepared by impregnating Cu-TCNQ NWs with ZnO colloid, then drying in vacuum [15]. Alejandro L. Briseno et al. also used the solution phase method to build two types of organic/ZnO heterojunction

Organic–Inorganic Heterojunction Nanowires

129

nanowires, including ZnO/P3HT[poly(3-hexylthiophene)] (Figure 6.1a) and ZnO/QT (didodecylquaterthiophene) (Figure 6.1b). As shown in Figure 6.1c and d, the ZnO/ P3HT and ZnO/QT core-shell heterojunction nanowires were successfully constructed with shell thicknesses of 7–20 nm and 6–13 nm, respectively. In terms of the ZnO/QT, every two QT molecules form π–π interactions with each other, and phosphonic acids

FIGURE 6.1  (a) Synthetic pathway for ZnO/P3HT nanowires; (b) Synthetic pathway for ZnO/QT nanowires; (c) Transmission electron microscopy (TEM) images of the ZnO/P3HT nanowires; (d) TEM images of the ZnO/QT nanowires. (Reprinted with permission from Ref 16. Copyright 2010 American Chemical Society)

130

Optoelectronic Organic–Inorganic Semiconductor Heterojunctions

group at the ends of different QT molecules form hydrogen bonds (P-O-H···O-P), and alkyl groups of different QT molecules have van der Waals force. These three intermolecular forces stabilize the structure of QT and align it vertically on the surface of ZnO. As for the ZnO/P3HT, P3HT molecules are arranged upside down on the surface of ZnO and there exist partially disordered layers [16]. P3HT:CdSe heterojunction nanowires were also fabricated on silicon by Xianfu Wang et al. Firstly, they prepared high-quality single crystals CdSe growing along [001] orientation with uniform diameter of 100 nm. Then, they mixed the P3HT solution with the CdSe solution and dropped it onto a silicon wafer for vacuum drying to get P3HT:CdSe heterojunction nanowires [17]. ZnO and polythiophene derivatives, (poly[2-(3-thienyl)-ethyloxy4-butylsulfonate)] (PTEBS) and poly[3-(potassium-6-hexanoate)thiophene-2,5-diyl] (P3KHT))-based core/shell coaxial nanowires, were also reported by Shanju Zhang and co-workers using electrostatic interactions to coat the organic layer on the ZnO nano-crystals [18]. Qing Yang et al. reported the fabrication of Zn/PEDOT:PSS coreshell nanowire and demonstrated effectively enhanced external efficiency of as-prepared ultraviolet light-emitting diode by piezo-phototronic effect [19]. 3D CoO/PPY coaxial nanowire array on nickel foam was also developed by Cheng Zhou et al. through template free solution process [20]. Hyunhyub Ko et al. reported hierarchical fibrillar arrays based on polycarbonate (PC) micropillar (μPLR) arrays decorated with ZnO nanowires (NWs) on flexible substrates. The as-prepared hierarchical PC μPLRs/ ZnO NWs exhibited excellent superhydrophobicity which was not observed from the PC μPLRs coated substrate and the blank substrate [21].

6.2.2 Template Method Combined with Electrochemical Polymerization Anodic aluminum oxide (AAO) is one of the most commonly used templates and is always used to grow organic–inorganic heterojunction nanowires assisted by electrochemical polymerization. The template method was firstly applied in the synthesis of organic–inorganic heterojunction nanowires by Sungho Park et al. [22]. They prepared three-segment Au-PPY (polypyrrole)-Au heterojunction via the template method associated with electrochemical polymerization of pyrrole. Besides, through analog method, they fabricated four-segment Au-PPY-Cd-Au heterojunction nanowires showing “diode” behavior [22]. Michal Lahav et al. synthesized core-shell and segmented organic– inorganic heterojunction nanowires via this method. By adjusting the pH of the solution (pH = 10.2), PANI (polyaniline)/Au core-shell structure was obtained by the template method. Then, PANI/Au segmented structure was also obtained by forming a self-assembled monolayer (SAM) of thioaniline on the top of Au [23]. Yanbing Guo et al. successfully fabricated CdS-PPY (polypyrrole) heterojunction nanowires using AAO template as shown in Figure 6.2a. The CdS-PPY nanowires were uniform, dense, and smooth with the diameter of 200–400 nm. Figure 6.2b obviously shows the interface between CdS and PPY, indicating the successful construction of heterojunction. As depicted in Figure 6.2c, which shows element distribution of C (green) and Cd (red), the end-to-end structure was revealed, suggesting the successful construction of heterojunction [24]. The schematic diagram of the synthesis of PBPB {Poly[1,4-bis(pyrrol-2-yl)benzene]}/CdS by AAO template is shown in Figure 6.2d. In sequence, PBPB and CdS were deposited into the AAO template

Organic–Inorganic Heterojunction Nanowires

131

FIGURE 6.2  (a) The side-view SEM image of CdS-PPY nanowires inside the AAO template; (b) The typical SEM image of a single CdS-PPY nanowire; (c) Element mapping of a single CdSPPY nanowire; (Reprinted with permission from Ref 24. Copyright 2008 American Chemical Society) (d) Synthesis procedure of PBPB/CdS heterojunction nanowire arrays; (Reprinted with permission from Ref 25. Copyright 2011 Royal Society of Chemistry) (e) Schematic illustration of the two different growing processes of PTh-CdS Core-Shell and Segmented Nanorod. (Reprinted with permission from Ref 26. Copyright 2009 American Chemical Society)

assisted by electrochemical polymerization. Finally, the AAO template was etched by NaOH and washed for several times to obtain PBPB/CdS heterojunction nanowires [25]. Besides, Yanbing Guo et al. explored two different growing processes of PTh(polythiophene)-CdS using AAO template method. As depicted in Figure 6.2e, at low current density, CdS would cover PTh nanowires to form core-shell structure (bottom-up growing). While at high current density, CdS would continue to grow at the top of PTh nanowires to form heterojunction (top-down growing) [26]. Haowei Lin et al. also fabricated PANI (polyaniline)/CdS via AAO template method. The diameter of single PANI/CdS nanowire is approximately 200 nm with the length of 10 μm [27]. Zheng Xue et al. prepared GD (graphdiyne)/CuS core-shell structure using AAO template method. Graphdiyne was firstly synthesized into AAO template via self-assembly process, then CuS was electrodeposited on the surface of the

132

Optoelectronic Organic–Inorganic Semiconductor Heterojunctions

graphdiyne to form GD/CuS heterojunction nanowire [28]. Giacomo Mariani et al. fabricated 3D nanostructured core-shell GaAs/PEDOT nanopillar arrays, which were fabricated via a fast, catalyst-free, bottom-up growth approach for GaAs and the following electrochemical deposition process for PEDOT coating [29]. Yuya Oaki et al. fabricated ZnO/PPY co-axial nanowire arrays on an indium tin oxide (ITO) substrate through low-temperature solution processes with irradiation of UV and visible light [30]. Recently, Xinhui Xia et al. demonstrated a few examples of metal oxide/­ conductive polymer co-axial nanowire including different nanoarray cores (nanowire and nanorod) of metal oxides (Co3O4 and TiO2) and different conducting polymer (PANI and PEDOT) shells through an electrochemical strategy [31]. 6.2.2.1 Template Method Combined with Pressure Injection The template method associated with pressure injection is another successful practice to prepare organic–inorganic heterojunction nanowires. For example, Yanbing Guo et al. applied AAO template to prepare CdS-OPV3 [oligo(p-phenylenevinylene)] heterojunction. OPV3 (Figure 6.3a) and CdS nanocrystals (Figure 6.3b) were firstly synthesized, then mixed to form colloid. A porous AAO membrane was immersed into the colloid and kept for 5–8 min under reduced pressure. Due to the differential pressure, the solution was immediately injected into the pores of the membrane. Finally, the AAO membrane was etched by NaOH solution and dried (Figure 6.3c). As depicted in Figure 6.3d, the diameter of CdS-OPV3 nanowires was 150–200 nm with the length of 2–5 μm. A clear interface between CdS and OPV3 was observed by HRTEM and is shown in Figure 6.3e, and the lattice spacing observed is about 0.204 nm, which is in good agreement with the plane spacing of the direction parallel (110) in wurtzite phase of CdS [32].

6.2.3 Vapor–Liquid–Solid Method The vapor–liquid–solid (VLS) method was used to grow GDY (graphdiyne)/ZnO organic–inorganic heterojunction nanowires by Xuemin Qian et al. for the first time. As shown in Figure 6.4a, they first grew ZnO nanowires on silicon substrate, then low molecular weight GDY vaporized and reacted with ZnO on the surface of ZnO to form metallic Zn. Small Zn molten droplets (melting point of Zn: 415 °C) formed on the tips of ZnO, then adsorbed GDY vapor. Subsequently, GDY molecules segregated to form GDY nuclei, and then GDY nanowires. During the growth, GDY pushed the Zn droplets up. Finally, Zn was oxidized to ZnO to form GDY/ZnO nanowires. TEM images (Figure 6.4b–d) showed the successful synthesis of the GDY/ZnO nanowires with uniform morphology [33].

6.3 THE APPLICATIONS OF ORGANIC–INORGANIC HETEROJUNCTION NANOWIRES The construction of organic–inorganic heterojunction nanowires not only retains the characteristics of each material but also produces novel electrical and catalytic properties through the interfacial interaction, which showed great potential in field emission, diode rectification, solar cells, photoelectric detection, and logic gates.

Organic–Inorganic Heterojunction Nanowires

133

FIGURE 6.3  (a) Molecular structure of oligo(p-phenylenevinylene) OPV3; (b) Transmission electron microscope (TEM) image of CdS nano-crystals; (c) Schematic illustration of the template synthesis of CdS-OPV3 hybrid nanorods; (d) TEM image of a few CdS-OPV3 nanorods. The inset in (d) is the selective area electron diffraction pattern (SAED) taken from the nanorods; (e) HRTEM image of CdS-OPV3 interface under a higher magnification. (Reprinted with permission from Ref 32. Copyright 2008 American Chemical Society)

6.3.1 Field Emission Field emission refers to the phenomenon that electrons are released from the surface of cathode under a strong electric field, which could be used in field-emission microscope, microwave power amplifier, and electron beam etch, etc. [11]. Huibiao Liu et al. loaded ZnO nanoparticles onto the Cu-TCNQ nanowires to form ZnO-CuTCNQ hybrid NWs. The turn-on field of ZnO-CuTCNQ (6.5 V/μm) was 3V lower than that of CuTCNQ (9.5 V/μm). They also synthesized In2O3-CuTCNQ

134

Optoelectronic Organic–Inorganic Semiconductor Heterojunctions

FIGURE 6.4  (a) Schematic illustration of the VLS process in the growth of GDY/ZnO nanowires; (b) Low magnification TEM image of GDY/ZnO nanowires; (c) and (d) Highmagnification TEM images of GDY/ZnO nanowires. (Reprinted with permission from Ref 33. Copyright 2012 Royal Society of Chemistry)

hybrid NWs by the same liquid phase method, which proved the universality of the method [15]. Yanbing Guo et al. prepared PTh-CdS core-shell and segmented structure using the AAO template method. Figure 6.5a shows that large-area and dense PTh-CdS core-shell was successfully fabricated. The half part of the PTh rounded by a shell of CdS is observed in Figure 6.5b. And the length of the shell can be controlled

Organic–Inorganic Heterojunction Nanowires

135

FIGURE 6.5  (a) Large area of PTh-CdS core-shell nanorod array; (b) Top view of PTh-CdS core-shell nanorod array under a higher magnification; (c) Side view of PTh-CdS core-shell nanorod array; (d) J-E plot of CdS nanorods, PTh nanorods, PTh-CdS segmented nanorods, and PTh-CdS core-shell nanorods; and (e) Corresponding F-N plot of those nanorods above. (Reprinted with permission from Ref 26. Copyright 2009 American Chemical Society)

by changing the time of electrodeposition. Figure 6.5c further indicates that the shell length of CdS was 2 μm. The field-emission performance of the PTh-CdS core-shell and segmented structure were tested and are shown in Figure 6.5d. The minimum turn-on field was achieved by the PTh-CdS core-shell structure (3.62 V/μm), which was lower than PTh (4.58 V/μm), the PTh-CdS segmented structure (6.72 V/μm) and CdS (10.41 V/μm). According to Figure 6.5e, the linear results showed that the field emission of the PTh-CdS core-shell and segmented structure was generated by the quantum tunneling mechanism [26]. To the PTh-CdS core-shell array, the triple

136

Optoelectronic Organic–Inorganic Semiconductor Heterojunctions

junctions are almost vertical to the anode, and electrons emitted from such triple junctions arrive at the anode easily, which resulted in the low turn-on field and low threshold field. However, as for the PTh-CdS segmented structure, triple junctions are almost parallel with the anode because the end-to-end structure, and the top fraction of CdS will shield the triple junction emissions which led to a higher turn-on field and threshold field. This proved that the structure of heterojunction can also regulate the field-emission performance. In summary, the field-emission performance of organic–inorganic heterojunction nanowires can be adjusted by changing the organic ligand, morphology, and heterostructure. Among them, triple junction formed by metal-semiconductor-vacuum and conductive polymer-carbon nanotube-vacuum plays an important role in the fieldemission properties (such as turn-on field).

6.3.2 Diode Rectification The diode (except for Schottky barrier diode) is mainly composed of P-N junction and can control the direction of the current in the circuit [11]. Diodes can be used for switching circuits, automotive lighting, light sources of electronic products, etc. [34]. Yanbing Guo et al. firstly prepared CdS-PPY organic–inorganic P-N junction nanowires and tested it as a diode rectification device. As shown in Figure 6.6c, TEM image showed the obvious interface, indicating that the CdS-PPY organic–inorganic P-N junction was successfully constructed. The diode performance of the CdS-PPY was measured (Figure 6.6d), and the light-response behavior of the CdS-PPY was similar to that of the PBPB/CdS mentioned above [14]. When light was introduced, the CdS-PPY showed strong light-response current, suggesting that it could be used as a potential diode material [24]. Haowei Lin et al. fabricated PTCM [poly(3-thiophene carboxylic acid methyl ester)]/PbS organic–inorganic heterojunction nanowires, which exhibited a rectification ratio of 15.7. Besides, distinct electrical switching properties were also observed for the PTCM/PbS heterojunction nanowires with a high ON/OFF ratio of 83.5 [35]. Nan Chen prepared and studied PbS/PPY heterojunction nanowires via the template method. Due to the intact, large-area contact between PbS and PPY, an obvious improvement was found for the PbS/PPY with a high rectification ratio (≥ 100). The PbS-PPY showed tight contact between organic and inorganic polymers and the larger heterogeneous junction area led to stronger interaction on the interface, which improved the property of diode rectification [36]. Then, Nan Chen et al. also utilized organic semiconductor polymer PBPB (P-type) and inorganic semiconductor CdS (N-type) to prepare PBPB/CdS organic–inorganic P-N junction nanowires and applied it in diode rectification (Figure 6.6a). As shown in Figure 6.6b, the diode rectification characteristics of PBPB/CdS heterojunction nanowires were investigated under different illumination conditions. In the dark, the light-response current of the PBPB/CdS was very low under the forward bias voltage. While white light was applied, the light-response signal increased obviously and improved with the increased light intensity. This phenomenon is due to the sensitivity of CdS to light and the tight contact with organic surface on the interface, resulting in a unique photoelectric response signal. Due to the different electron affinity between CdS and PBPB, the electrons and holes generated in the photoexcitation process

Organic–Inorganic Heterojunction Nanowires

137

FIGURE 6.6  (a) The PBPB/CdS heterojunction nanowire arrays between two electrodes under an applied electric field; (b) Typical current–voltage (I-V) curves for PBPB/CdS heterojunction nanowire arrays under light illumination with different intensities; (Reprinted with permission from Ref 25. Copyright 2011 Royal Society of Chemistry) (c) Typical TEM image of a single CdS-PPY heterojunction nanowire; (d) Typical current-voltage (I-V) curves for a single CdS-PPY heterojunction nanowire under light illumination with different intensities at room temperature (inset: SEM image and EDS line analysis of the measured nanodevice). (Reprinted with permission from Ref 24. Copyright 2008 American Chemical Society)

would be transferred through CdS and PBPB, respectively. As the light intensity increases, the number of electrons and holes increases, so the diode effect is better when the light intensity is stronger [25]. Based on this interesting phenomenon, they believed that the photoelectric response of CdS must be selective. Therefore, they further fabricated PANI/CdS nanowires, and the light-response behavior of the PANI/ CdS was tested under different wavelength illumination (254–610 nm). They found that the light-response current of the PANI/CdS irradiated with a blue light of 420 nm was much greater than that of the sample irradiated with other wavelengths, demonstrating the selectivity of the PANI/CdS. Besides, they proposed that the PANI/CdS can be used as a photoelectric detector to detect the intensity of blue light [27]. To summarize, organic–inorganic P-N junction nanowires can be constructed by selecting suitable N-type and P-type semiconductor materials, including organic

138

Optoelectronic Organic–Inorganic Semiconductor Heterojunctions

semiconductor polymer and inorganic semiconductor. Excellent light-controlled diode rectification performance can be realized via the use of organic–inorganic heterojunction nanowires. However, diode performance, such as the rectifier ratio and the simplification of preparation process, needs to be further improved.

6.3.3 Solar Cells Solar cells are electronic devices that absorb sunlight and convert it into electricity, making them more environmentally friendly than conventional batteries and charge– discharge batteries. Due to its low cost, simple preparation and high stability, organic–inorganic hybrid thin-film solar cells have attracted much attention [11]. Alejandro L. Briseno et al. fabricated ZnO/P3HT and ZnO/QT organic–inorganic heterojunction nanowires and tested their photovoltaic performance. Figure 6.7a shows a schematic diagram of the photovoltaic performance test of a single ZnO/ oligothiophene nanowire. The SEM image and the typical I-V curve of the ZnO/ P3HT are shown in Figure 6.7b. The ZnO/P3HT produced a short circuit (Jsc) of 0.32 mA/cm2, an open circuit voltage (Voc) of 0.40 V, and an efficiency of 0.036%. Similar to the ZnO/P3HT, the ZnO/QT nanowires yield a Jsc of 0.29 mA/cm2, a Voc of 0.35 V, and an efficiency of 0.033% (Figure 6.7c). They proposed that single-nanowire devices improved the shunt resistance by removing the short-circuit path in bulk ZnO array devices [16]. Yanbing Guo et al. utilized a single CdS-PPY organic–inorganic P-N junction nanowire to convert light energy into electricity. The construction of heterojunction significantly enhanced the range and intensity of light absorption, which is beneficial for the working condition device under visible light. The CdSPPY delivered an efficiency of 0.018% under low light intensity of 6.05 mW/cm2, while photovoltaic characteristics were not found in individual CdS and PPY nanowires. Additional tests were carried out on individual CdS and individual PPY nanowires to demonstrate that the photovoltaic properties are caused by unique organic/ inorganic p-n junctions formed in a single nanowire [37]. Recently, Sanghoon Yoo et al. also reported the fabrication of Au-PPY-CdSe axial nanorod and its application as a solar cell. They found that unadorned Au-PPY-CdSe-Au nanorods with up to 1.1% power conversion efficiency could be obtained by using a porous Au nanorod electrode in the core of the PPY-CdSe nanorod. Due to the presence of electrophilic nitrogen atoms in the pyrrole ring unit, PPY can be chemically adsorbed to Au to generate tight contact. Nitrogen atom allowed electrons to jump from the p-conjugated molecular chain to the Au, thereby reducing band bending at the Au-PPY interface. Thus, the dissociation efficiency of excitons was enhanced [38]. Hao Xin et al. reported that the poly(3-butylthiophene) nanowires/[6,6]-phenyl-C61-butyric acid methyl ester (P3BT-nw/C61-PCBM) nanocomposite solar cell with P3BT nanowire (8–10 nm in width and up to 5–10 μm in length) as the donor component embedded in a sea of C61-PCBM acceptors showed a dramatically improved energy conversion efficiency compared with its counterpart P3BT:C61-PCBM blend. It was concluded that the substantially better hole transport ability of P3BT nanowire network in the P3BT-nw/C61-PCBM solar cells contributed to the improved performance [39]. Yajie Zhang and co-workers reported organic single-crystalline P-N junction nanoribbons of CuPc and F16CuPc, which showed a relatively lower energy conversion efficiency (η) of ∼0.007% under 100 mW/cm2 light intensity [40].

Organic–Inorganic Heterojunction Nanowires

139

FIGURE 6.7  (a) Schematic configuration of a discrete ZnO/oligothiophene nanowire solar cell; (b) SEM image of a discrete ZnO/P3HT nanowire device fabricated by EBL and the corresponding current-voltage characteristics; (c) SEM image of a discrete ZnO/QT nanowire device and the current-voltage characteristics. The devices were measured under AM 1.5 irradiation (100 mW/cm2). (Reprinted with permission from Ref 16. Copyright 2010 American Chemical Society)

140

Optoelectronic Organic–Inorganic Semiconductor Heterojunctions

In summary, organic–inorganic heterojunction nanowires showed certain photovoltaic performance, proving their feasibility as solar cell material. But their photovoltaic performance, including the short circuit, the open circuit voltage, and the efficiency, needs to be improved.

6.3.4 Photoelectric Detection Owing to incident radiation, the conductivity of semiconductors would change a lot leading to a light-dependent behavior. This phenomenon can be used for photoelectric detection which has wide applications, such as optical communication, sensors [17]. Guozhen Shen et al. prepared P3HT:CdSe heterojunction nanowire photodetectors showing a improved photocurrent with a short recovery time (≤0.1 s). The photocurrent would enhance with the light wavelength (350–650 nm) increased, while it would decrease with the light wavelength (700–850 nm) increased. This demonstrated a strong light-dependent behavior of the P3HT:CdSe heterojunction nanowires. The high hole transport rate of P3HT, the high conductivity of CdSe, and the synergistic absorption spectra of each component in the visible spectrum were responsible for the enhanced optical response and stability of the P3HT:CdSe [17]. Haowei Lin et al. successfully fabricated uniform PANI/CdS nanowire photodetectors as shown in Figure 6.8a and b. The electrical properties of the PANI/CdS were measured as shown in Figure 6.8c. The experimental results (Figure 6.8d) showed that the PANI/CdS was more sensitive to blue light (420 nm) than other tested wavelengths. The rectification ratio of the PANI/ CdS improved (from 11.4 to 34.1) with enhanced light intensity (Figure 6.8e). Besides, as depicted in Figure 6.8f, the PANI/CdS exhibited excellent stability under the illumination of 420 nm. Some unique properties of the PANI/CdS were thought to be beneficial for high response speeds: (1) firstly, the CdS had high crystallinity and the trap density caused by defects was greatly increased; (2) secondly, high surface-to-volume ratio of the PANI/CdS allows surface defects and dangling bonds act as recombination centers, enhancing the recombination of carriers and thus reducing the decay time; (3) finally, the lowering of the recombination barrier is caused by Fermi level pinning [27]. Then, Haowei Lin et al. demonstrated the feasibility of CdS/PPV (p-phenylene vinylene) as a photodetector. The results showed that the photocurrent of the CdS/PPV enhanced dramatically under the illumination of 545 nm while the photocurrent of the CdS was almost 0. The significant enhancement of the photoelectric performance of CdS/PPV hybrid nanowire arrays is mainly due to the matching energy level between PPV and CdS. PPV can be used as a bridge to consume or transfer photogenerated holes and electrons, therefore restraining their recombination [41].

6.3.5 Logic Gates Logic gates are the basic components in an integrated circuit that perform logic operations through high and low potentials. In general, a logic gate carries out logical operations on one or more inputs and produces an output [42]. Nan Chen et al. firstly reported and proposed a novel principle device for logic gates fabricated by using a three-segment (organic–inorganic–organic) heterojunction nanowire and the synthesis process is shown in Figure 6.9a. The component of

Organic–Inorganic Heterojunction Nanowires

141

FIGURE 6.8  (a) Cross-view SEM image of PANI/CdS heterojunction nanowires; (b) Highresolution SEM image of PANI/CdS heterojunction nanowires; (c) Working model of PANI/ CdS heterojunction nanowire array device; (d) Typical I-V curves of PANI/CdS heterojunction nanowire arrays in dark and under illumination of different wavelength light; (e) Irradiance dependence of the rectification ratio of PANI/CdS heterojunction nanowire arrays in the dark and under 420 nm light illumination; (f) On/off switching of PANI/CdS heterojunction nanowire arrays upon pulsed illumination from 420 nm wavelength light with a power density of 5.21 mW/cm2. (Reprinted with permission from Ref 27. Copyright 2011 American Chemical Society)

142

Optoelectronic Organic–Inorganic Semiconductor Heterojunctions

FIGURE 6.9  (a) Schematic diagram of the synthesis process of the EPP heterojunction nanowire arrays; (b) SEM image of a single EPP heterojunction nanowire device made by focus ion beam; (c) Standard symbols for logic OR2 gate (two-input OR gate); (d) The signal output of the logic two-input OR gate constructed using EPP nanowire; and (e) The output data of the logic two-input OR gate constructed using EPP nanowire. (Reprinted with permission from Ref 42. Copyright 2013 Springer Nature)

three-segment heterojunction nanowire was poly(3,4-ethylenedioxythiophene) (PEDOT), PbS, and polypyrrole (PPY), respectively. The device was prepared by EPP (PEDOT/PbS/PPY) nanowires via ion beam (Figure 6.9b). Figure 6.9c showed a typical equivalent circuit (logic OR2 gate) which was consistent with the behavior of the EPP nanowires. The experimental results (Figure 6.9d) revealed a “logical

Organic–Inorganic Heterojunction Nanowires

143

gates” behavior: when the input was higher than “1” (voltage ≥ 5 V), there was an output; while the input was lower than “1,” there was no output. Detailed data results are summarized in Figure 6.9e [42].

6.4 SUMMARY AND PERSPECTIVE This chapter introduces the research progress of organic–inorganic heterojunction nanomaterials. Organic–inorganic heterojunction nanowires have been successfully fabricated by four methods (solution phase method, template method combined with electrochemical polymerization, template method combined with pressure injection, and VLS method) and showed excellent performance in field emission, diode rectification, solar cells, photoelectric detection, and logic gates. Among these researchers, Yuliang Li et al. extensively prepared and studied all kinds of organic–inorganic heterojunction nanowires, explored the growth mechanism of them, and explained the structure–activity relationship between structure and performance. To sum up, the interface structure of organic–inorganic heterojunction nanowires can be adjusted by changing the organic ligand, growth mode, and morphology to adjust the electrical and catalytic properties. In the future, novel synthesis methods (such as VLS method) need to be proposed and practiced to extend the materials diversity of organic–inorganic heterojunction nanowires. For the materials reported so far, they may also be used in new electronic devices (such as logic gates). At present, organic–inorganic heterojunction nanowires have been widely studied and applied in electric and photoelectric field. However, novel physical or chemical properties brought by the construction of heterojunction need to be further explored. For example, Yanbing Guo et al. fabricated Cu-TCNQ nanowires showing excellent performance for CO catalytic oxidation for the first time, which opens up a novel application field [43]. Further improvement would be achieved if Cu-TCNQ nanowires can be further compounded with other metals (such as Pt) or metal oxides (such as CuO) to form organic–inorganic heterojunction nanowires. It is a reminder that organic–inorganic heterojunction nanowires still have a lot of potential of application to be discovered. On the other hand, the conductivity of the material would be greatly improved for organic–inorganic heterojunction nanowires under the illumination or the dark [15,17]. This suggests that organic– inorganic heterojunction nanowires synthesized by reasonable design may be applied in electrochemical water splitting or photoelectrochemical water splitting.

REFERENCES 1. Meng, F. E. I., Morin, S. A., Forticaux, A., et al. Screw dislocation driven growth of nanomaterials. Acc. Chem. Res., 2013, 46, 1616–1626. 2. Zang, L., Che, Y., Moore, J. S. One-dimensional self-assembly of planar π-conjugated molecules: adaptable building blocks for organic nanodevices. Acc. Chem. Res., 2008, 41, 1596–1608. 3. Mei, J., Diao, Y., Appleton, A. L., et al. Integrated materials design of organic semiconductors for field-effect transistors. J. Am. Chem. Soc., 2013, 135, 6724–6746. 4. Figueira-Duarte, T. M., Mullen, K. Pyrene-based materials for organic electronics. Chem. Rev., 2011, 111, 7260–7314.

144

Optoelectronic Organic–Inorganic Semiconductor Heterojunctions

5. Virkar, A. A., Mannsfeld, S., Bao, Z., et al. Organic semiconductor growth and morphology considerations for organic thin-film transistors. Adv. Mater., 2010, 22, 3857–3875. 6. Schwartz, G., Tee, B. C. K., Mei, J., et al. Flexible polymer transistors with high pressure sensitivity for application in electronic skin and health monitoring. Nat. Commun., 2013, 4, 1–8. 7. Facchetti, A. Organic semiconductors: Made to order. Nat. Mater., 2013, 12, 598–600. 8. Pramod, P., Thomas, K. G., George, M. V. Organic nanomaterials: morphological control for charge stabilization and charge transport. Chemistry-An Asian Journal, 2009, 4, 806–823. 9. Henson, Z. B., Müllen, K., Bazan, G. C. Design strategies for organic semiconductors beyond the molecular formula. Nat. Chem., 2012, 4, 699–704. 10. An, B. K., Gierschner, J., Park, S. Y. π-Conjugated cyanostilbene derivatives: a unique self-assembly motif for molecular nanostructures with enhanced emission and transport. Acc. Chem. Res., 2012, 45, 544–554. 11. Qian, X., Liu H., Li, Y. Self-assembly low dimensional inorganic/organic heterojunction nanomaterials. Chin Sci Bull, 2013, 58, 2686–2697. 12. Wang, Z. Characterizing the structure and properties of individual wire-like nanoentities. Adv. Mater., 2000, 12, 1295–1298. 13. Guo, Y., Xu, L., Liu, H., et al. Self-assembly of functional molecules into 1D crystalline nanostructures. Adv. Mater., 2015, 27, 985–1013. 14. Zheng, H., Li, Y., Liu, H., et al. Construction of heterostructure materials toward functionality. Chem. Soc. Rev., 2011, 40, 4506–4524. 15. Liu, H., Cui, S., Guo, Y., et al. Fabrication of large-area hybrid nanowires arrays as novel field emitters. J. Mater. Chem., 2009, 19, 1031–1036. 16. Briseno, A. L., Holcombe, T. W., Boukai, A. I., et al. Oligo- and Polythiophene/ZnO hybrid nanowire solar cells. Nano Lett., 2010, 10, 334–340. 17. Wang, X., Song, W., Liu, B., et al. High-performance organic-inorganic hybrid photodetectors based on P3HT:CdSe nanowire heterojunctions on rigid and flexible substrates. Adv. Funct. Mater., 2013, 23, 1202–1209. 18. Zhang, S., Pelligra, C. I., Keskar, G., et al. Directed self-assembly of hybrid oxide/polymer core/shell nanowires with transport optimized morphology for photovoltaics. Adv. Mater., 2012, 24, 82–87. 19. Yang, Q., Liu, Y., Pan, C., et al. Largely enhanced efficiency in ZnO nanowire/p-polymer hybridized inorganic/organic ultraviolet light-emitting diode by piezo-phototronic effect. Nano Lett., 2013, 13, 607–613. 20. Zhou, C., Zhang, Y., Li, Y., et al. Construction of high-capacitance 3D CoO@ polypyrrole nanowire array electrode for aqueous asymmetric supercapacitor. Nano Lett., 2013, 13, 2078–2085. 21. Ko, H., Zhang, Z., Takei, K., et al. Hierarchical polymer micropillar arrays decorated with ZnO nanowires. Nanotechnology, 2010, 21, 295305. 22. Park, S., Chung, S. W., Mirkin, C. A. Hybrid organic-inorganic, rod-shaped nanoresistors and diodes. J. Am. Chem. Soc., 2004, 126, 11772–11773. 23. Lahav, M., Weiss, E. A., Xu, Q., et al. Core-shell and Segmented polymer-metal composite nanostructures. Nano Lett., 2006, 6, 2166–2171. 24. Guo, Y., Tang, Q., Liu, H., et al. Light-controlled organic/inorganic P-N junction nanowires. J. Am. Chem. Soc., 2008, 130, 9198–9199. 25. Chen, N., Qian, X., Lin, H., et al. Synthesis and characterization of axial heterojunction inorganic–organic semiconductor nanowire arrays. Dalton Trans., 2011, 40, 10804–10808. 26. Guo, Y., Liu, H., Li, Y., et al. Controlled core-shell structure for efficiently enhancing field-emission properties of organic-inorganic hybrid nanorods. J. Phys. Chem. C, 2009, 113, 12669–12673.

Organic–Inorganic Heterojunction Nanowires

145

27. Lin, H., Liu, H., Qian, X., et al. Constructing a blue light photodetector on inorganic/ organic p-n heterojunction nanowire arrays. Inorg. Chem., 2011, 50, 7749–7753. 28. Xue, Z., Yang, H., Gao, J., et al. Controlling the interface areas of organic/inorganic semiconductor heterojunction nanowires for high-performance diodes. ACS Appl. Mater. Interfaces, 2016, 8, 21563–21569. 29. Mariani, G., Wang, Y., Wong, P. S., et al. Three-dimensional core-shell hybrid solar cells via controlled in situ materials engineering. Nano Lett., 2012, 12, 3581–3586. 30. Oaki, Y., Oki, T., Imai, H. Enhanced photoconductive properties of a simple composite coaxial nanostructure of zinc oxide and polypyrrole. J. Mater. Chem., 2012, 22(39), 21195–21200. 31. Xia, X., Chao, D., Qi, X., et al. Controllable growth of conducting polymers shell for constructing high-quality organic/inorganic core/shell nanostructures and their opticalelectrochemical properties. Nano Lett., 2013, 13, 4562–4568. 32. Guo, Y., Li, Y., Xu, J., et al. Fabrication of homogeneous hybrid nanorod of organic/ inorganic semiconductor materials. J. Phys. Chem. C, 2008, 112, 8223–8228. 33. Qian, X., Ning, Z., Li, Y., et al. Construction of graphdiyne nanowires with high-conductivity and mobility. Dalton Trans., 2012, 41, 730–733. 34. Zhao, J., Li, S., Tao, M. The working principle and main application of the diode. Public Commun. Sci. Technol., 2010, 10, 45–46. 35. Lin, H., Liu, H., Qian, X., et al. Synthesizing axial inserting p-n heterojunction nanowire arrays for realizing synergistic performance. Inorg. Chem., 2013, 52(12), 6969–6974. 36. Chen, N., Xue, Z., Yang, H., et al. Growth of axial nested P-N heterojunction nanowires for high performance diodes. Phys. Chem. Chem. Phys., 2015, 17, 1785–1789. 37. Guo, Y., Zhang, Y., Liu, H., et al. Assembled organic/inorganic p-n junction interface and photovoltaic cell on a single nanowire. J. Phys. Chem. Lett., 2010, 1, 327–330. 38. Yoo, S. H., Liu, L., Ku, T. W., et al. Single inorganic-organic hybrid photovoltaic nanorod. Appl. Phys. Lett., 2013, 103, 1–4. 39. Xin, H., Kim, F. S., Jenekhe, S. A. Highly efficient solar cells based on poly (3-butylthiophene) nanowires. J. Am. Chem. Soc., 2008, 130, 5424–5425. 40. Zhang, Y., Dong, H., Tang, Q., et al. Organic single-crystalline p-n junction nanoribbons. J. Am. Chem. Soc., 2010, 132, 11580–11584. 41. Lin, H., Chen, K., Li, M., et al. Constructing a green light photodetector on inorganic/ organic semiconductor homogeneous hybrid nanowire arrays with remarkably enhanced photoelectric response. ACS Appl. Mater. Interfaces, 2019, 11, 10146–10152. 42. Chen, N., Chen, S., Ouyang, C., et al. Electronic logic gates from three-segment nanowires featuring two p–n heterojunctions. NPG Asia Mater., 2013, 5, 59. 43. Hu, S., Xiao, W., Yang, W., et al. Molecular O2 activation over Cu(I)-mediated C≡N bond for low temperature CO oxidation. ACS Appl. Mater. Interfaces, 2018, 10, 17167–17174.

7

Electroluminescence of Organic Molecular Junction in Scanning Tunneling Microscope Xiaoguang Li

CONTENTS 7.1 Introduction: Molecular Junctions and Devices............................................ 147 7.2 Transport Mechanism in Molecular Junctions............................................... 149 7.2.1 Coherent Transport��������������������������������������������������������������������������� 150 7.2.2 Incoherent Transport������������������������������������������������������������������������� 152 7.3 Optical Properties of Molecular Junctions.................................................... 154 7.4 Special Phenomena: Hot Luminescence and Upconversion.......................... 155 7.4.1 Hot Luminescence���������������������������������������������������������������������������� 156 7.4.2 Upconversion Electroluminescence�������������������������������������������������� 157 7.5 Summary and Outlook................................................................................... 161 References............................................................................................................... 161

7.1 INTRODUCTION: MOLECULAR JUNCTIONS AND DEVICES In this chapter, we focus on the optoelectronic properties of molecular devices. Over the past few decades, advancements in nanofabrication techniques and quantum theory of electronic transport and optics have enabled us to explore and to understand the basic characteristics of rudimentary electronic circuits in which single molecules or molecular assemblies are used as functional building blocks. The molecular devices discussed here should not be confused with organic electronics, where molecular materials are investigated as possible components of various macroscopic electronic devices. Compared with conventional functional devices composed of bulk semiconductor materials or even 2D semiconductors, a single molecule or molecular assembly is naturally nanoscale and can achieve multiple functions due to its rich and mutually coupled internal degrees of freedom. The study of single-molecule devices is now no longer limited to the early electronic transport properties [1,2], but involves a series of novel physical phenomena, including mechanics, thermoelectricity, optoelectronics, and spintronics, as depicted in Figure 7.1a [3]. Interesting new phenomena have been found to emerge due to competition between those different transition processes 147

148 Optoelectronic Organic–Inorganic Semiconductor Heterojunctions

FIGURE 7.1  (a) Schematic of the electronic transport, mechanic response, thermoelectric phenomena, optical effects, and spin-dependent transport of single-molecule junctions. (Source: Aradhya and Venkataraman 2013 [3]) (b) Schematic junction geometry with multimonolayer stacking and localized electrical excitation from a nanotip. (Source: Dong et al. 2010 [5]) (c) Schematic of the STM-induced single-molecule emission from a single H2Pc ­molecule on two monolayer NaCl. (Source: Chen et al. 2019 [8])

Electroluminescence of Organic Molecular Junction in Scanning

149

with different energy and time scales. These findings are highlighting single-­molecule devices broad application prospects and important scientific significance. Molecular luminescence is an elemental and crucial process that forms the basis of various organic optoelectronic devices [4]. Despite decades of in-depth research, many fundamentally important issues remain unresolved, partly because the experimental probes available in this field are limited. It has been demonstrated recently that the molecular luminescence induced by scanning tunneling microscope (STM) provides an unprecedented opportunity to explore many important optoelectronic phenomena, including hot luminescence [5–7], upconversion electroluminescence [5,8], Fano resonance [9–11], superradiance [12], and Electrofluorochromism [13]. The typical experimental setup for those observations is shown in Figure 7.1b and c. The observed phenomena are not only fundamentally intriguing but also may play an important role in optimizing the performance of optoelectronic devices. STM utilizes the sensitivity of the tunneling rate of electrons to the width of the potential barrier between metal tip and substrate and can obtain extremely high spatial resolution. When molecules are inserted into STM, the appearance of the molecular electronic orbitals could drastically change the transport characteristics of the tunneling electrons. In turn, the optical properties of the molecule itself could be greatly modulated by the STM due to the surface plasmon mode of the nanocavity formed by metal tip and substrate. The most interesting thing is that the metal nanostructure of STM could help to solve the issue that light-matter coupling is too weak for single-molecule devices due to the huge size mismatch between the wavelength of visible light and the size of single molecules. Consequently, the STM system with molecular junction can combine the high resolution of space, time, and energy spectra, which can help us systematically analyze the observed optoelectronic phenomena [14–17]. In this chapter, we will focus on two peculiar electroluminescence phenomena in metal-molecule-metal junctions formed by a molecule inserted in STM: the hot luminescence and upconversion electroluminescence in Section 7.4. To elucidate the underlying physics of both phenomena, we will focus on some necessary fundamental knowledge about transport mechanisms of the metal-molecule-metal junction in Section 7.2, and the optical properties of the molecule and the metal nanocavity formed in an STM in Section 7.3. As we will gradually introduce in the following sections, our main emphasis for metal-molecule-metal junctions is on the energylevel alignment, energy scales between different pieces, and time scales of different processes in the system. Unfortunately, this chapter does not provide a comprehensive review of optoelectronics in a molecular junction. Readers are encouraged to look at the other chapters of this book as well as the many excellent reviews and treatises that are now available on this topic [18–25].

7.2 TRANSPORT MECHANISM IN MOLECULAR JUNCTIONS In this section, we discuss the transport mechanism in the metal-molecule-metal junction, covering both the coherent and incoherent transports [26]. We first give an example of coherent transport, the so-called resonant tunneling model, where the electron tunnels through the molecule by forming a virtual charged molecule state.

150

Optoelectronic Organic–Inorganic Semiconductor Heterojunctions

Then we will introduce the incoherent transport, which is important for the electroluminescence. The discussion is only based on simple toy models or handwaving arguments, trying to convey the core physical ideas instead of the tedious calculation details.

7.2.1 Coherent Transport Different from the macroscopic electronic conduction, where the resistance mainly comes from the electron–phonon interaction, the conductance G of the atomic scale STM junction should be described by the Landauer formula as G

e2 h

T , i

(7.1)

i

where e2/h is the quantum unit of the conductance determined by electron charge e and Planck constant h, and Ti is the transmission of an individual transition mode. Essentially, the quantum nature of electrons should be considered in such a small spatial distance, and the conductance of the system is closely related to the potential barrier tunneling problem in quantum mechanics, namely, the resistance mainly comes from the interface scattering. When a molecule is inserted into the STM metal–metal junction, the transmission of an electron can be evaluated by using Breit–Wigner formula as T  E ,V  

4 S T

, 2 2  E   V      S  T 

(7.2)

where ΓS (ΓT) indicates the coupling between the substrate (tip) and molecule, and E and ϵ(V) are the energies of tunneling electrons and molecular electronic orbitals, respectively, as shown in Figure 7.2a. A tunneling process of an electron (hole) through this metal-molecule-metal junction is depicted by the red dot (circle) and arrows in Figure 7.2b. When the Fermi surface of the tip is higher than that of the substrate, an electron can tunnel through the molecule LUMO (Lowest Unoccupied Molecular Orbital) with a probability determined by ΓT, and then a virtual/transient (or real as long as E > ϵ) charged molecular anion state M* is formed, and finally the electron in the LUMO of state M* can tunnel to the substrate with a probability determined by ΓS. Conversely, we can also consider a hole state transport through the molecule by forming a transient cation state with a hole in HOMO (Highest Occupied Molecular Orbital). Both processes contribute a positive-charge current from the substrate to tip. So, the coherent transport through molecular junctions is mainly determined by the strength of the metal-molecule coupling as well as by the alignment between the molecular electronic orbital (usually HOMO and LUMO) and the metal Fermi level. Another way to understand this tunneling process is that due to the hybridization of the molecular orbitals and the electronic state in metal, the molecular orbitals acquire a finite broadening that provides a nonzero density of state aligning with the energy of the tunneling electron.

Electroluminescence of Organic Molecular Junction in Scanning

151

FIGURE 7.2  (a) Schematic of energy-level alignment and molecule–electrode coupling in STM system. (b) Schematic of the resonant tunneling process through the HOMO and LUMO of the molecule. (c) Current and dI/dV vs. bias voltage in the resonant tunneling model for ΓT = 0.2 eV, ΓS = 0.01, and ϵ0 = 0.1.

By further obtaining a quantitative result for the current-bias relation in the STM junction, the tunneling current I at a bias voltage V can be evaluated by adopting the Landauer formula as where

2e 2 I V   h



 dE T  E,V   f  E,V

S

  f (E,VT )



n f  E,    E   / kBT e  1

(7.3) (7.4)

152

Optoelectronic Organic–Inorganic Semiconductor Heterojunctions

is the state occupation at the energy E for the metal electrode with the chemical potential μ, density of state n, temperature T, and Boltzmann constant kB. For the transmission T(E, V), it could be assumed that  V    0  V



T ,   S  T 

(7.5)

indicating the bias voltage dependence of the molecular orbital with its energy measured from the Fermi level of the metal substrate. Clearly, the voltage drop is considered only to happen at the tunneling barrier between the two metal electrodes and the molecule. In the STM system discussed here, we have ΓS ≫ ΓT. The expression for ϵ(V) simply reflects the fact that if one of the coupling strength is much greater than the other, the molecular orbital follows the shift of the chemical potential of the electrode that is better coupled. Figure 7.2b shows the current and dI/dV under different bias evaluated from the Landauer formula (Equation 7.3). The result clearly exhibits the so-called Coulomb blockade transport behavior of the molecular junction. Essentially, at a low-bias voltage and weak-coupling regime, the electron tunneling rate is extremely small because tunneling electrons cannot “classically” run into the molecule. In this case, the Fermi levels of the electrodes lie somewhere within the HOMO–LUMO gap of the molecule and are far from the HOMO and LUMO levels. The transmission of the electron is low, and the increase of the current with bias is due to the increase of the available tunneling electrons in the bias window between the Fermi level of the tip and substrate. As the bias further increases until the Fermi level of the tip approaches the molecular LUMO level, the conductance drastically climbs up due to the increasing transmission rate. At the resonant condition with E − ϵ(V) = 0, we obtain the largest 4 S  T transmission 2 . As the bias further increases with the LUMO entering

  S  T 

into the bias window, the conductance of the molecular junction jumps to a higher value with one quantum unit increasing. In this case, the tunneling electron can really run into the molecule and stay at the LUMO level, and the original high-order tunneling process relying on the two simultaneous tunneling events could now be done by two independent so-called charge injection processes. The tunneling rate is thus largely increased. By now, we have seen that the transport property of the metal-molecule-metal junction is mainly determined by the coupling between the molecule and metal electrodes, and the energy-level alignment between molecular orbital and Fermi level of metal. The different parameters of the model, including the energy levels and the coupling constants, could be considered as phenomenological parameters, while they could in principle be obtained from a fit to the experimental results or from theoretical ab initio calculation.

7.2.2 Incoherent Transport Electroluminescence typically originates from the generation of electron–hole pairs by incoherent transport through the metal-molecule-metal junction. In this subsection, we present an introductory description of incoherent transport.

Electroluminescence of Organic Molecular Junction in Scanning

153

Probably, the simplest picture of the incoherent transport can be understood when both HOMO and LUMO levels lie in the bias window. In this case, as shown in Figure 7.3a, after one electron from the tip tunneling to the molecular LUMO, the next tunneling procedure could be another electron in HOMO tunneling from the molecule to the substrate, instead of the corresponding coherent case with the electron further tunneling from LUMO to the substrate. So, the whole incoherent process includes two different charge injection steps and leaves the molecule at a neutral yet excited state. Here, we explicitly emphasize on the “neutral”, because the transport process involves the ground and excited states of the molecules not only for the neutral species but also for the cation and anion states. For example, if we consider the molecule as a two-level electronic system with level 0 occupied and level 1 empty for the neutral ground state, i.e., HOMO, the charge injection for the neutral molecule requires the Fermi level of the tip rising to be higher than LUMO. However, for a cation state with only one electron at level 0, the charge injection happens when the Fermi level of the tip reaches the spin-singlet state S1 as shown in Figure 7.3a. So, for a charge injection process, the net charge of the molecule is varied, and the position of the molecular levels participating in the carrier injection is different from what is referred to in the one-electron molecular orbital picture. With the above understanding in mind, you may immediately realize that in the above-mentioned incoherent transport, after the electron tunneling into LUMO, the

FIGURE 7.3  (a) Schematic of different electronic state configurations of a single molecule. (b) Schematic of inelastic tunneling processes.

154

Optoelectronic Organic–Inorganic Semiconductor Heterojunctions

second charge injection from the molecule to substrate does not require the “neutral” HOMO higher than the Fermi level of the substrate. Actually, since the molecule is now at the anion state, the second charge injection only requires that the Fermi level of the substrate lies ET1 below the LUMO level, and then the injection is energetically allowed. Or conversely, as the bias increases, if the HOMO becomes aligned with the Fermi level of the substrate, one electron at level 0 can then tunnel into the substrate and leave the molecule at a cation state. The subsequent charge injection from the tip to molecule can happen as long as the corresponding Fermi level is higher than the T1 energy level. Here, it is worth to emphasize that as long as the spin of the tunneling electrons of the two charge injection processes is different, the spin state of the molecule can be changed, namely, both the triplet and singlet states can be excited by the inelastic transport. This is clearly different from the excitation by an external light source. Another possible situation is that the bias voltage could reach the excitation energy of the T1 state with neither HOMO nor LUMO aligning with the Fermi level of metal electrodes, as shown in Figure 7.3b. In this case, the two-step charge injection transport cannot happen. However, similar as in the coherent transport, the excitation could happen mediated by a virtual cation state, with one electron tunneling from the tip to molecular HOMO, and simultaneously one electron tunneling from the molecular LUMO to the substrate. This process can be considered as electron– electron scattering, which may also be described as shown in Figure 7.3c with an electron directly tunneling from the tip to substrate and scattering the electron in the molecule to a higher energy level. These two processes cannot be distinguished as long as the spin state of the molecule is not changed. So far, we have introduced the fundamental pictures of coherent and incoherent transports in the metal-molecule-metal junction. It would be very challenged to calculate the occurrence probability of different processes in actual systems, but the understanding of the above pictures can help us to explain the observed electroluminescence phenomena by combining the rate equation method and the different transition rates obtained from experimental fitting.

7.3 OPTICAL PROPERTIES OF MOLECULAR JUNCTIONS Different from the conventional semiconductor functional devices, in single-molecule devices, the light-emission property of the metal-molecule-metal junction depends not only on the intrinsic electronic structures of the molecule but also on its surrounding electromagnetic environment dominated by the nanocavity, formed by the metallic tip and substrate. This nanocavity plays an important role in the optical response of the junction due to the corresponding surface plasmon mode. In metal materials, surface plasmons describe the collective oscillation mode of electrons at the surface of the material, and their corresponding highly localized enhanced electric fields give surface plasmons a wide range of applications [27–43]. Extremely strong electromagnetic near-field allows surface plasmon to be used as a nanoantenna to improve the coupling of various optoelectronic devices with the external light field and thus enhance absorption and emission [32,36,38,43]. In addition, the highly localized characteristics also allow it to break the diffraction limit of the optical field, which is essential for the miniaturization of optoelectronic devices.

Electroluminescence of Organic Molecular Junction in Scanning

155

In a vacuum, the emission of excited molecules due to the spontaneous radiation is a pure quantum phenomenon, which originates from the interaction between the transition dipole of the molecule and the vacuum fluctuations of the electromagnetic field. It is well known that the presence of dielectric nanocavity surrounding the molecule could modify the photon density of states, and therefore change the spontaneous decay rate, known as the Purcell effect. The strength of the Purcell effect is described by the Purcell factor 3



F

6 Q , 2 V  2 

(7.6)

where Q = ω/Γ with resonant frequency ω and linewidth Γ is the quality factor of the nanocavity, V is the volume of the cavity, and λ is the wavelength of the emitted light. For a dielectric cavity, the diffraction limit gives V > (λ/2)3, and therefore conduct the upper limit for the Purcell factor F  62 Q [44].  A metallic plasmonic nanocavity can also change the photon density of states and thus utilize the Purcell effect to dramatically enhance the spontaneous decay rates of the molecule. Since the surface plasmon can break the diffraction limit of the optical field, the fundamental limitation for its Purcell factor comes from the quality factor Q, more specifically, the linewidth Γ of the nanocavity. Essentially, a large part of the radiative energy transferring from the molecule to metal nanocavity will be dissipated through the non-radiative decay in metals due to the electron–electron scattering. In addition, the de-excitation of the molecule close to metals also suffers from the dissipation due to the possible charge transfer and non-radiative energy transfer, so the distance between the molecule and metal tip and substrate should be carefully controlled. Different insulating films, such as Al2O3, NaCl, and several layers of molecules themselves, have been used as spacers to avoid those issues, as shown in Figure 7.1b and c. An example of the role of the metal nanocavity in the de-excitation of the molecule is shown in Figure 7.4, where the total and radiative decay rates of a molecule between two spherical metal nanoparticles are evaluated by using the generalized Mie theory [6]. In comparison with the result in a vacuum, the radiative decay rate of the molecule is increased by 3–5 orders of magnitude in optical wavelengths, driving the radiative lifetime for a nanosecond to picosecond regime. By comparing the radiative and total decay rates, we also see the presence of a large portion of non-­radiative decay. In addition, we can find that the gold dimer exhibits a large non-radiative decay rate in the high-frequency region, which is due to the interband transition loss in gold above ~2 eV. Therefore, to achieve a large radiative enhancement for the molecular junction, we need to carefully select proper plasmonic materials and control the size and shape for the tip and substrate in the STM.

7.4 SPECIAL PHENOMENA: HOT LUMINESCENCE AND UPCONVERSION As we have mentioned in the previous two sections, the energy alignment largely determines the transport behaviors, while the coupling between the metallic nanocavity formed by the tip and substrate strongly affects the emission of the molecule.

156

Optoelectronic Organic–Inorganic Semiconductor Heterojunctions

FIGURE 7.4  (a) Schematic picture of the system setup. The tunneling junction is modeled with two metal spheres and the molecule is treated as an electric dipole locating on the axis of the dimer. (b) Radiative and total decay rates for an electric dipole placed in a vacuum, or within a gold dimer and silver dimer, respectively. (Source: Chen et al. [6])

When we bring the transport and optics together in this metal-molecule-metal junction, some representative electroluminescence phenomena can be observed, including the hot luminescence [5–7], upconversion [5,8], and the Fano phenomenon [9–11], superradiance [12], and Electrofluorochromism [13]. In this section, we discuss two of them, i.e., hot luminescence and upconversion electroluminescence of the molecular junction in an STM. The main emphasis is on the competition between different transport and optical processes due to their different energy and time scales.

7.4.1 Hot Luminescence The excitation of the molecule usually combines both the electronic and vibrational parts, namely, the electron jumps to the high energy state and atoms start to oscillate. The corresponding excited state can be denoted as |e, v⟩. In a vacuum, light emission from an excited molecule follows Kasha’s rule, which states that only emission from the lowest vibrational level of the excited state is possible. This is because, for free molecules, the vibrational damping rate is typically 1012 s−1, while the spontaneous emission rate determined by the transition dipole is typically 107 ∼ 108 s−1. Consequently, the molecule will first cool down to the vibrational ground state |e, 0⟩, and then make a radiative transition to electronic ground state |g, v⟩. The emission

Electroluminescence of Organic Molecular Junction in Scanning

157

final state may not be |g, v  =  0⟩, because the equilibrium position of a vibrational mode upon the excited electronic state is usually not coincident with that of the ground electronic state. The Kasha’s rule, however, is founded to be broken in some molecular junctions in a STM. In a work done by Dong et al. [5], people find the modulation of molecular emission profile as well as the presence of high energy emission, which seems coming from the directly radiative transition from a “hot” electronic excited state |e, v⟩ to ground state |g, v′⟩ without going through a picosecond cooling process. A schematic plot of the excitation and emission processes in the experiment has been shown in Figure 7.5a. The molecule is excited by a charge injection process with one electron tunneling from |g, 0⟩ (HOMO) to drain and another electron tunneling from source to higher energy level in the molecule, exciting both electronic and vibrational motions. The normal luminescence starts from a cooling process labeled by γvib, following by the transition labeled by γs from |e, 0⟩ to the ground state, while the hot luminescence labeled by γh will not go through the cooling process. Within the framework of the quantum master equation proposed earlier by Johansson et al., and by considering the spontaneous emission enhancement due to the metallic nanocavity, Chen et al. [6] show that the electroluminescence spectral profile of the metal-molecule-metal junction can be largely modulated. A hot luminescence could be observed, when the radiative decay rate of the molecule is enhanced to be comparable with the vibration relaxation rates. To obtain the luminescence spectra, two essential quantities of the molecule, the radiative decay rate and the state population at the dynamic equilibrium, should be evaluated. An example of the enhancement of the radiative decay rate has been shown in Figure 7.4 in Section 7.3. Here, the results of the spectral profiles at different system configurations have been shown in Figure 7.5, which shows that the spontaneous decay rate of the molecule can be strongly modulated by tuning the geometry and material of the nanocavity. Figure 7.5b shows that the spectra can be selectively modulated depending on the quantum efficiency of the cavity, which is related to the geometrical structure of the nanocavity. In turn, the intrinsic properties of the molecule will also greatly affect the luminescence spectrum, as shown in Figure 7.5d and e, in which the intensity and direction of the transition dipole can strongly affect the thermal luminescence peak with an energy higher than 2 eV. In essence, the presence of hot luminescence depends on the delicate competition between the different transition processes. As shown in Figure 7.5c, the theoretical study also indicates that some hot luminescence results may be overlooked before, because the corresponding peak could overlap with a norm luminescence peak.

7.4.2 Upconversion Electroluminescence Upconversion electroluminescence in an STM is a good example to show how the energy and time scales of different transition processes and the energy-level alignment affect the optoelectronic properties of a system. Upconversion luminescence is a conceptually counterintuitive quantum phenomenon in which a single emitted photon has higher energy than that of the excitation sources. In the STM system, it means that the energy of an emitted photon is higher than energy loss by a single tunneling

158

Optoelectronic Organic–Inorganic Semiconductor Heterojunctions

FIGURE 7.5  (a) Schematic diagram of energy levels and various transition processes within the system formed by a molecule in a tunnel junction. (b) Molecular fluorescence spectra in vacuum and in a tunnel junction for gold dimers. (c) Illustration of contributions from different channels for molecular emission. (d) Molecular fluorescence spectra in the junction of a gold dimer for different dipole moment and gap distance. (e) Molecular fluorescence spectra and quantum efficiency for electric dipoles parallel and perpendicular to the axis of the gold dimer. (Source: Chen et al. [6])

Electroluminescence of Organic Molecular Junction in Scanning

159

electron, provided by the bias voltage. Clearly, this phenomenon requires a multielectron excitation process of fluorescent radiation. Therefore, all the different mechanisms proposed in earlier studies of upconversion electroluminescence, such as intermolecular triplet–triplet annihilation (TTA) [45,46] and molecular vibrationassisted plasmonic pumping [5,47], involve intermediate states to capture and store multiples of the energy quanta from the excitation source. In a recent experiment done by Chen et al. [8], it has been shown that the upconversion electroluminescence could happen at the single-molecule and low-tunneling current level. Figure 7.1c exhibits the geometrical configuration of the system with isolated H2Pc molecules between tip and substrate, and electronically decoupled by a two-monolayer thick NaCl spacer. The main experiment findings include two aspects. One is the upconversion phenomenon, which is interpreted to be mediated by the spin-triplet excited state in the single molecule. The other is the observation of the three distinct efficiency-bias regions, implying the different transport and excitation mechanisms. The conclusion is based on the detailed analysis of the energy and time scales of different processes in the system, and also the energy alignment between the molecule and substrate. Considering the experiment setup with the low tunneling current, the observed upconversion luminescence cannot be explained by most of the previously proposed intermediate states, including the surface plasmon of the nanocavity and vibrational mode of the molecule. In this experiment, the tunneling current is as low as ~100 pA, corresponding to an average time interval of ~1 ns between the electron-tunneling events. However, the surface plasmon of the nanocavity has a typical femtosecond lifetime [48], and the molecular vibrational mode has a typical picosecond lifetime [49]. Consequently, the spin-triplet state of a free H2Pc molecule with a lifetime ~100 μs [50] seems the only possible intermediate state in this system. The bias voltage threshold for the upconversion also implies the spin-triplet state mediating the upconversion. Figure 7.6a shows the optical spectra from the single molecule at different negative bias voltages. The normal electroluminescence happens at Vb = –2.5 V with the energy of a tunneling electron higher than that of the spin-singlet state at ~1.81 eV. The upconversion electroluminescence could be observed at Vb = –1.7 V with a clear emission peak at 1.81 eV. The voltage threshold for the upconversion can be seen in Figure 7.6b at around –1.2 V, which coincides with the energy of the spin-triplet state of H2Pc molecule. As shown in Figure 7.6c, the emission intensity shows three distinct stepwise increases in different voltage regions with constant current, implying the different excitation efficiency or say different inelastic excitation rate in three bias regions: (I) Vb